Sei sulla pagina 1di 9

Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Optimization of carboxyl-functionalized mesoporous silica for the selective T


adsorption of dysprosium

Takamasa Kanekoa,b, Fukue Nagatab, Shinichi Kugimiyaa, Katsuya Katob,
a
Materials Chemistry Course, Graduate School of Engineering, Aichi Institute of Technology, 1247 Yachigusa, Yakusa-cho, Toyota, 470-0392, Japan
b
National Institute of Advance Industrial Science and Technology, 2266-78, Anagahora, Moriyamaku, Nagoya, Aichi, 463-8560, Japan

A R T I C LE I N FO A B S T R A C T

Keywords: Dysprosium (Dy), one of the rare-earth elements, was adsorbed on the four types of mesoporous silica, having
Mesoporous pore size 3–22 nm and functionalized with amino and carboxyl groups on the surface. The influence of the pore
Silica size on the selective Dy ion adsorption capacity was examined using a metal ion mixture. The synthesized
Carboxyl materials was characterized by field emission scanning electron microscopy, transmission electron microscopy,
Dysprosium
Brunauer-Emmett-Teller analyzer. In addition, the functionalized materials were characterized by thermo-
Selectivity
gravimetric analyzer, Fourier transform infrared, and zeta-potential. The adsorption amount of Dy ions on the
functionalized particles was evaluated by inductively coupled plasma optical emission spectrometry. The ma-
terial having a large pore size (22 nm) as well as amino and carboxyl chains manifested an excellent adsorption
capacity for Dy ions (32.9 μg/mg). The recycling efficiency of adsorption-desorption process retained approxi-
mately 100% of the capability at least five cycles. These results suggested that the functionalized material with
pore size 22 nm can be used for an adsorbent for the selective and efficient removal of Dy ions from mixtures.

1. Introduction 0.146 mmol/g [9,10]. Several studies have also reported on Dy ion
adsorption using the functionalized amino and carboxyl groups on the
Dysprosium (Dy), one of the rare-earth elements as neodymium solid surface. In terms of efficiency, silica samples with high surface
magnets, is useful in the automobile industry. Although the increasing area such as porous particles are promising materials since they have
of Dy consumption in recent years is attributed to the increasing de- high adsorption capacity [11–13]. Therefore we assumed that meso-
mand for its factor, however, it is difficult to assure a constant supply porous silica (MPS) was considerably attractive for developing highly
because of low production and high cost. In 2013, approximately 90% efficient materials for Dy adsorption.
of the total world production of Dy element occurred in China. Ordered MPS has uniform nanosize pores that are formed with the
However, the significant problem related to this element is generated aid of surfactant micelles [14–20]. MPS is an inorganic material with
due to the amount of supply and price stability since the most part of carefully controlled pore sizes (2–50 nm), large surface area, and high
resources was dependent on other countries. Therefore, usual and pore volume [20–22]. We previously reported that enzyme im-
simple recycling systems of Dy are necessary. mobilization on MPS showed enhanced enzyme activity and stability
To date, various materials were reported as adsorbents for metal [23]. For immobilization inside the pores, the available free space must
ions, e.g., ceramic materials and polymers [1–4]. It is well known that be considered to provide sufficient enzyme mobility for retaining their
various adsorbates were adsorbed on the surface of these inorganic catalytic activity. Additionally, the free space ensures substrate diffu-
materials via physisorption; however, they manifested low selective sion to the catalytic sites. Herein, we demonstrate that the porous
adsorption [5,6]. The removal of a particular ion from the mixture of materials with adequate pore structure allow for the adsorption of
various metal ions is an important industrial approach [7,8]. In pre- metal ions and focus on the effects of Dy adsorption amounts on the
vious study, Dy ions were adsorbed from a metal ion mixture on non- functionalized MPS with different pore sizes. MPS with pore size and
porous the silica gel modified with the amino and carboxyl groups. It pore distribution, which were easily controlled, was synthesized via a
was reported that these silica materials selectively adsorbed Dy on their templating process using a surfactant as the organic template and tet-
surface based on a complex formation between the organic groups and raethoxysilane (TEOS) as silica sources. The formation of the solid
the outermost shell of the Dy ion, with an adsorption capacity of supports is based on the hydrolysis and polycondensation processes of


Corresponding author.
E-mail addresses: takamasa-kaneko@aist.go.jp (T. Kaneko), katsuya-kato@aist.go.jp (K. Kato).

https://doi.org/10.1016/j.jece.2018.09.018
Received 20 July 2018; Received in revised form 31 August 2018; Accepted 14 September 2018
Available online 18 September 2018
2213-3437/ © 2018 Elsevier Ltd. All rights reserved.
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

Scheme 1. Schematics of amino-functionalized mesoporous silica (MPS) materials employing (a) APTES and −NH2 and (b) N6-APTES and −2NH2 and the following
carboxyl-functionalized MPS materials employing diglycolic anhydride: (c) −NH-COOH and (d) −2NH-2COOH, respectively.

Fig. 1. FE-SEM images of (a) mesoporous silica (MPS) (3 nm), (b) MPS (5 nm), (c) MPS (12 nm), and (d) MPS (22 nm).

silica precursors. The process is that allows designing characteristics the 2. Experimental procedure
synthesized materials. MPS has SiOH groups on the surface, allowing
for functionalization with various organic groups. Herein, we synthe- 2.1. Materials
sized MPS with various pore sizes (3, 5, 12, and 22 nm) using a sol–gel
method, while the surface was functionalized with the amino and car- 1-Hexadecytrimethylammoniumchloride (CTAC) and Pluronic P123
boxyl groups using commonly employed coupling reagents. The amount (P123) were purchased from Sigma-Aldrich Co., St. Louis., MO. TEOS
of metal ion adsorbed on MPS was systematically monitored. and 3-aminopropyltrimethoxysilane (APTES) were commercially ob-
In addition, the effect of the pore size on the Dy ion adsorption tained from Shin-Etsu Chemical Co., Tokyo, Japan. Diglycolic hydride
capacity of the resulting materials was studied. The recycling property was purchased from Tokyo Chemical Industry Co., Tokyo, Japan. 1,3,5-
of the adsorbents was also investigated. Trimethylbenzene (TMB), acetic acid, toluene, n-decane, aqueous am-
monia, ethanol, and hydrochloric acid were obtained from Wako Pure
Chemical Industries, Osaka, Japan. N-(6-aminohexyl)aminopropyl-
trimethoxysilane (N6-APTES) was obtained from AZmax Co., Chiba,
Japan.

5991
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

Fig. 2. TEM images of (a) MPS (3 nm), (b) MPS (5 nm), (c) MPS (12 nm), and (d) MPS (22 nm).

2.2. Characterization 2.3. Synthesis of MPS

The morphology of the synthesized MPS materials was investigated 2.3.1. Synthesis of MPS (3 and 5 nm) particles
using field emission scanning electron microscopy (FE-SEM; S-4300, MPS particles (3 and 5 nm) were synthesized according to a pre-
Shimadzu Co., Kyoto, Japan) at an accelerating voltage of 10.0 kV, via viously reported synthesis procedure and using CTAC as an organic
transmission electron microscopy (TEM; JEM 2010, JELO Ltd., Tokyo, template [21]. CTAC (2.6 g) was added to 30 mL of aqueous solution
Japan) at an accelerating voltage of 200 kV. Scanning TEM (STEM; (pH 0.5), and the mixture was stirred for 2 h at room temperature. To
JEOL JEM-2100 Plus, JOEL Ltd., Tokyo, Japan) images were operated obtain mesopores of 5 nm, TMB (2.6 g) was added to the solution
at 200 kV, and a copper grid with coating carbon was used for metal containing CTAC. Subsequently, TEOS (3.5 g) was added dropwise and
substrates. The nitrogen (N2) adsorption and desorption measurements the solution was stirring at 4 °C. After 2 h stirring, NH3 (3.0 g, 28 wt.%,
were performed using a Shimadzu TriStar 3000 system after preheating 14.7 M) was added to the solution and the reaction mixture was stirred
at 105℃, and the specific surface area was calculated according to the for 24 h. Finally, the product was centrifuged at 6000 rpm for 20 min.
Brunauer–Emmett–Teller (BET) method, and the pore size distribution washed and calcined at 500 °C for 5 h in air (heating rate of 1℃/min) to
was obtained by the Barrett–Joyner–Halenda (BJH) method. The remove the organic template.
amounts of the amino and carboxyl functional groups on MPS were
established in heating temperature range from room temperature to
1000℃ (10℃/min) using a thermogravimetric analyzer (TG-DTA; TGA- 2.3.2. Synthesis of MPS (12 and 22 nm) particles
502, Shimadzu Co., Kyoto, Japan). The organic functional groups on The procedure for preparing MPS (12 nm) is described as follows:
the surface were also analyzed via Fourier transform infrared (FT-IR) first, triblock copolymer P123 (1 g), used as the template, was com-
spectroscopy in the wavenumber range of 400–4000 cm−1 using an pletely dissolved in HCl (35 mL, 1.07 M). Then, n-decane (7.6 g) was
MFT-2000 spectrophotometer (JASCO Co., Tokyo, Japan). The spectra added to the mixture. After 2 h stirring, ammonium fluoride (NH4F;
were obtained in the transparent mode using a pellet made of 99% KBr 11.5 mg) was added to the solution. Next, TEOS (2.13 g) was added to
and 1% sample. Additionally, to prove the existence of functional the mixture and the mixture was stirred for 1 h. Subsequently, the re-
groups on MPS, the surface potential of MPS was measured using a zeta- action mixture was incubated at 40 °C for 20 h in a temperature-con-
potential analyzer (ELSZ-1000, Otsuka Electronics Co., Tokyo, Japan). trolled bath, after which the suspension was hydrothermally treated for
Before the measurements, the particles were dispersed in distilled water 24 h at 110 °C. The MPS (22 nm) particle was also prepared using TMB
with sonication for 5 min. The amount of metal ions adsorbed on MPS (1 g) as a pore expander. First, P123 was dissolved in distilled water
was determined using inductively coupled plasma optical emission (26 mL). After gentle stirring, TMB (1 g) was added to the mixture and
spectrometry (ICP-OES; IRIS Advantage, Thermo Fisher Scientific, MA). the mixture was stirred overnight. Next, NH4F (11.5 mg) and TEOS
(2.1 g) were added to the solution and the reaction mixture was stirred
for 10 min at room temperature. Then, a concentrated HCl (4 mL) so-
lution was gradually added and the solution was stirred for 30 min at

5992
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

Fig. 3. N2 adsorption–desorption (A–D) isotherms and pore size distribution curves (a–d) of the MPS materials. (A, a) MPS (3 nm), (B, b) MPS (5 nm), (C, c) MPS
(12 nm), and (D, d) MPS (22 nm). These samples were measured after the samples were preheated at 105℃.

Table 1 room temperature. Subsequently, the resulting mixture was incubated


The amount of amino and carboxyl groups on functionalized MPS materials. at 35 °C for 20 h, after which it was transferred to an autoclave and
Sample -NH2a -NH-COOHa −2NH2a −2NH-2COOHa treated at 110 °C for 24 h. To remove the template, MPS was calcined at
(μg/m2)b (μg/m²)b (μg/m²)b (μg/m²)b 500 °C for 5 h. The heating temperature were ranged from room tem-
perature to 500℃ (heating rate 1 ℃/min).
MPS (3 nm) 20 137 26 106
MPS (5 nm) 234 412 349 494
MPS (12 nm) 377 637 292 709
2.4. Surface functionalization of MPS
MPS (22 nm) 293 629 379 797

a
Amino-functionalized MPS materials using APTES; −NH2 or N6-APTES; To adsorb metal ions on the MPS surface, amino- and carboxyl-
−2NH2, followed by functionalization with carboxyl groups; −NH-COOH and functionalized MPS materials were prepared using a method similar to
−2NH-2COOH, respectively. that described in previous research [24,25]. Amino-functionalized MPS
b
The amount of functional groups on MPS materials was calculated from the was prepared by dispersing MPS (100 mg) in toluene (10 mL), and 1 mL
weight losses in TG curves. of 3-aminopropyltriethoxysilane (APTES; −NH2) or N-(6-aminohexyl)
aminopropyltrimethoxysilane (N6-APTES; −2NH2) were added to the
solution for 10 min. Next, the mixture was refluxed for 6 h. The product
was centrifuged at 6000 rpm for 20 min, and the precipitate was

5993
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

6000 rpm for 20 min. The supernatant was filtered using a 0.45 μm
filter, and the amount of metal ions in supernatant was measured via
ICP-OES. The amount of metal ions adsorbed on carboxyl-functiona-
lized MPS was calculated from the change in the concentration of the
Dy ions into the solution using Eq. (1). Where Qb and Qa is the initial
Dy ions concentration and after adsorption, respectively, and I0 is the
initial Dy ions content.

Eq = (Qb – Qa) I0 (1)

2.6. Adsorption isotherms

Adsorption isotherm experiments were conducted at room tem-


Fig. 4. FTIR spectra of MPS materials. a) red line denotes MPS (12 nm). b) The perature using MPS (12 nm)-2NH-2COOH and MPS (22 nm)-2NH-
MPS (12 nm)-NH2 and c) MPS (12 nm)-2NH2 samples are denoted by yellow 2COOH. These samples having large amount of Dy ions adsorption were
and blue lines, respectively. d) The MPS (12 nm)-NH-COOH and e) MPS added to Dy ions solutions with various concentrations (2–10 ppm), and
(12 nm)-2NH-2COOH samples are denoted by green and purple lines, respec- adsorption process was conducted for 2 h at the room temperature.
tively. These sample were also analyzed in the wavenumber range of After centrifugation, the Dy ions concentration of the supernatant was
400–4000 cm−1. measured by ICP-OES, and adsorption amount of Dy ions on functio-
nalized MPSs was calculated. To establish the metal ion adsorption
mechanism, the data were further processed by Langmuir and
Freundlich equations. Detailed definitions of the variables can be found
in the literature [26–29]. The Langmuir adsorption isotherm model is
given by Eq. (2):

Qe = (QmKLC)/ (1 + KLC) (2)

where Qe (μg/mg) is the amount of metal ions on MPS adsorbed at


equilibrium, C (mol/L) is the equilibrium concentration in the super-
natant, and KL is the Langmuir constant. Qm denotes the maximum
monolayer surface coverage adsorption capacity on MPS.
The Freundlich isotherm model indicates multilayer adsorption, and
the equation of this model is expressed as follows:

Qe = KFCe1/n (3)

where KF (mg/g) is the Freundlich constant that denotes the adsorption


capacity, the 1/n value stands for the favorable adsorption condition.

Fig. 5. TEM image of MPS (22 nm) −2NH-2COOH. 2.7. Cycling tests

Cycling tests was conducted using carboxyl-functionalized MPS


washed with ethanol and acetone, and dried at room temperature.
(22 nm)-2NH-2COOH. The adsorption and desorption experiments
Carboxyl-functionalized MPS was prepared as follows: 150 mg of MPS-
were performed at room temperature only for Dy ions. The adsorption
NH2 and MPS-2NH2 were redispersed into toluene (15 mL), and the
process was carried out as follows: first, carboxyl-functionalized MPS
solution was stirred at 50 °C for 10 min. Subsequently, diglycolic an-
materials (5 mg) were added to 25 mL of Dy ions solution (10 ppm) and
hydride (261 mg) was added to the solution and the mixture was in-
the mixture was stirred for 2 h. The supernatant was separated by
cubated at 50 °C. After 1 h, the reaction mixture was stirred at room
centrifugation at 6000 rpm for 20 min, and the solution was filtered
temperature. Finally, the products were washed with ethanol and
using a 0.45 μm filter. To desorb Dy ions from the adsorbent, the pre-
acetone for dryness (Scheme 1).
cipitate was added to 12.5 mL of 1 M HNO3 solution and the mixture
was stirred for 2 h. The supernatant was separated by centrifugation at
2.5. Adsorption of metal ions on functionalized MPS materials 6000 rpm for 20 min, and the solution was filtered using a 0.45 μm
filter. Subsequently, the adsorbent was washed twice with distilled
The concentration of metal ions (copper, Cu; dysprosium, Dy; iron, water. The adsorption experiment was repeated five times. After each
Fe; neodymium, Nd; and zinc, Zn) was 10 ppm for each element. The filtration, the amount of Dy ions in the supernatant was measured via
solutions were prepared in distilled water. The adsorption tests were ICP-OES. The amount of Dy ions adsorbed on carboxyl-functionalized
conducted in a solution at pH 1.75, adjusted by titration with 1 M MPS was calculated from the change in the concentrations of Dy ions in
HNO3. Metal ion (Cu, Dy, Fe, Nd, and Zn) mixture were used in this the solution before adsorption; Qc and after desorption; Qd. To evaluate
experiment, and the adsorption experiments were carried out at the the cycling property of MPS, the adsorption capacity QR was calculated
room temperature (20℃). The adsorption process was carried out as based on the concentration of Dy ions in the supernatant using Eq. (4).
follows: first, carboxyl-functionalized MPS (2 mg) was added to 10 mL The value of 100% capacity illustrated the adsorption Dy ions desorbed
of the metal ion solution (Cu, Dy, Fe, Nd, and Zn; concentration was from the MPS completely by desorption process.
10 ppm for each element) and the mixture was stirred for 2 h at room
temperature. The supernatant was separated by centrifugation at QR (%) = (Qd/Qc) × 100 (4)

5994
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

Fig. 6. EDX profiles and EDX mapping images


of MPS (22 nm) −2NH-2COOH. STEM ele-
mental mapping images of MPS
(22 nm)−2NH-2COOH. Yellow, violet, and
green indicate silicon, nitrogen, and oxygen,
respectively. STEM images were operated at
200 kV, and a copper grid with coating carbon
was used for metal substrates.

Fig. 7. Adsorption of the five metal ions (Cu, Dy, Fe, Nd, and Zn) on functionalized MPS materials, with initial concentration of 10 ppm, pH: 1.75, room temperature
and adsorption time of 2 h.

3. Results and discussion isotherms are summarized, and the values for the specific surface area
of MPS with different pore sizes (3, 5, 12, and 22 nm) were 998, 865,
3.1. Properties of synthesized MPS particles 608, and 546 m2/g, respectively, while those for the pore volumes were
1.27, 0.73, 1.44, and 1.95 cm3/g, respectively. The actual size of the
The as-synthesized MPS materials with various pore sizes were pores was controlled at approximately 2.6, 5.2, 12.5, and 22.5 nm.
characterized using FE-SEM, TEM, and N2 physisorption. The mor- Based on these values, the samples were denoted as MPS (3, 5, 12, and
phology of the synthesized MPS materials was characterized using FE- 22 nm), respectively. From these results, it could be concluded that the
SEM and TEM. The FE-SEM and TEM images of the synthesized MPS synthesized MPS materials have a mesoporous structure and optional
particles (3, 5, 12, and 22 nm) are shown in Fig. 1(a)–(d) and Fig. 2, pore diameter.
respectively. According to the FE-SEM images in Fig. 1(a)–(d), MPS
(3 nm) and MPS (5 nm) was composed of small particles (50–500 nm), 3.2. Characterization of functionalized MPS materials
while the MPS (12 nm) and MPS (22 nm) show agglomerated
morphologies with polyhedral particles with mainly in the range of MPS particles were functionalized with APTES and N6-APTES, fol-
1–10 μm. As observed in TEM images, MPS particles was confirmed the lowed by functionalization with carboxyl groups using diglycolic an-
hexagonal structure and exhibited the presence of mesoporous channels hydride, as shown in Scheme 1, and named as (a) −NH2, (b) −2NH2,
with each diameter and appearance in Fig. 2(a)–(d). (c) −NH-COOH, and (d) −2NH-2COOH, respectively.
N2 physisorption was performed for MPS particles, and the resulting To confirm the functionalization of the MPS surface, FTIR, TG-DTA,
adsorption–desorption isotherms [(A)–(D)] and pore size distribution N2 physisorption, and ζ-potential analyses were performed. To de-
curves [(a)–(d)] are shown in Fig. 3. MPS with pore size (3, 5, 12, termine the amount of the amino and carboxyl groups on the MPS
22 nm) were exhibited the isotherms of type IV, furthermore, MPS surface, thermal gravimetric analysis was performed. The corre-
having pore size 12 nm and 22 nm are indicative of H1 hysteresis loops, sponding weight losses were attributed to the decomposition of organic
respectively. The textural properties calculated based on the obtained groups on the surface. From the exothermic peak attributed to the

5995
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

797 μg/m2, respectively. According to these results, the weight losses


were higher, with each successive functionalization indicating that the
amount of the amino and carboxyl groups on MPS progressively in-
creased with the pore size. In particular, as the amount of organic
groups on MPS-2NH-2COOH increased in comparison with that on
MPS-NH-COOH, it was possible to observe the highly increasing
amount of functionalized carboxyl groups on MPS (12 and 22 nm)-2NH-
2COOH at 709 and 797 μg/m2, respectively. The amount of organic
groups on MPS (12 and 22 nm)-2NH-2COOH greatly increased in
comparison to that on MPS (3 and 5 nm)-2NH-2COOH. It can be as-
certained that in the last two cases, the pores of MPS were obstructed by
the organic groups, while for MPS (12 and 22 nm), the pores were
sufficiently large to avoid obstruction during the second step of func-
tionalization.
ζ-Potential measurements were performed to analyze the surface
charge for each functionalized MPS. It was observed that the obtained
values were mainly affected by the functional groups attached on the
surface. The ζ-potentials MPS (22 nm), MPS (22 nm)-NH2, and MPS
(22 nm)-2NH-2COOH were 8.83, 18.1, and −37.4 (mV), respectively.
From these results, the presence of the amino and carboxyl functional
groups on MPS-2NH2 and MPS-2NH-2COOH was confirmed. The FTIR
Fig. 8. Illustration of pore size and surface functionalization on MPS with small spectra of MPS samples are shown in Fig. 4. The spectra of all MPS
(3 nm) and large (12 nm) pore sizes. (A) Functionalization of MPS with amino
samples contain absorption bands characteristic to the silanol groups
groups, (B) subsequent functionalization of MPS (pore size = 3 nm) with car-
O–Si–O and Si–O–Si at 943 and 1089 cm−1, respectively. For the
boxyl groups, and (C) subsequent functionalization of MPS (pore size = 12 nm)
amino-functionalized MPS samples, including MPS (12 nm)-NH2 and
with carboxyl groups.
MPS (12 nm)-2NH2, the absorption bands at 2865, 2980 and
1540 cm−1 attributed to −CH2, −CH2 and NeH bonds, respectively,
Table 2 were noticed, providing evidence for the presence of organic groups on
Langmuir and Freundlich constants for Dy ion adsorption. the silica surface [30–32]. Thus, according to the registered spectra for
Langmuir plot Freundlich plot MPS (12 nm)-NH2, MPS (12 nm)-2NH2, MPS (12 nm)-NH-COOH, and
MPS (12 nm)-2NH-2COOH, the amino groups derived from APTES
Sample Qma R2
1/nb R2
(−NH2) or N6-APTES (−2NH2) were bonded to the surface. Finally, for
MPS (12 nm) 39.1 0.99 0.54 0.95 the carboxyl-functionalized MPS samples, including MPS (12 nm)-NH-
−2NH-2COOH COOH and MPS (12 nm)-2NH-2COOH, the absorption peak at
MPS (22 nm) 44.8 0.99 0.59 0.94 1680 cm−1 corresponding to the eC]O group was observed for both
−2NH-2COOH samples [32]. Therefore, the FTIR spectra of the MPS (22 nm)-NH2 and
a MPS (22 nm)-2NH2 samples showed the presence of amino groups,
Qm denotes the maximum monolayer surface coverage adsorption capacity
on MPS materials. while those of the MPS (22 nm)-NH-COOH and MPS (22 nm)-2NH-
b
1/n value stands for the favorable adsorption condition. 2COOH samples proved the presence of both the amino and carboxyl
groups. In addition, the TEM images of the structure and morphology of
MPS materials before and after functionalization were shown in Fig. 5.
As shown in the TEM images, it was indicated that the porous structure
was retained like small domains of cylindrical pores. The results dis-
played no difference in the ordered mesoporous structure between the
samples before and after functionalization, thereby showing that the
porosity of the functionalized MPS materials was preserved. To confirm
the presence of functionalized groups in nanoparticles, the elemental
distribution of MPS (22 nm)-2NH-2COOH was investigated by STEM-
EDX analysis. EDX spectrum of the functionalized MPS (22 nm)-2NH-
2COOH is depicted in Fig. 6, which indicates the presence of silicon,
oxygen, and nitrogen. The nitrogen is based on the functionalized
groups containing amino groups. The elemental mapping confirms a
uniform distribution of the functionalized groups around the nano-pore
structure of silica nanoparticles.
Fig. 9. Cycling test of MPS (22 nm)-2NH-2COOH performed five times for Dy
ions adsorption, with initial concentration of 10 ppm, pH: 1.75, room tem-
3.3. Metal ion adsorption on MPS samples
perature and adsorption time of 2 h. The desorption process was performed by
adding to the 1 M HNO3 solution and the mixture was stirred for 2 h.
Subsequently, the absorbent was washed twice with distilled water. The ad- Different metal ions (i.e., Cu, Dy, Fe, Nd, and Zn) were adsorbed on
sorption-desorption experiment was repeated five times. the amino- and carboxyl-functionalized MPS materials via physical
adsorption. The amount of ions adsorbed at equilibrium on these
samples is shown in Fig. 7. As shown in the figure, the adsorption of Cu,
decomposition of the amino and carboxyl groups, the amount of or-
Fe, Nd, and Zn ions on the functionalized MPS samples did not occur,
ganic groups on the MPS samples were calculated (Table 1). Therefore,
whereas the Dy ions were found to have adsorbed. According to pre-
the amount of organic groups on MPS (3, 5, 12, and 22 nm)-NH-COOH
vious research, the adsorption of Dy ions has been explained via a
was 137, 412, 637, and 629 μg/m2, respectively, while on MPS (3, 5,
complexation reaction occurring between the Dy ions and the carbonyl
12, and 22 nm)-2NH-2COOH, the amount was 106, 494, 709, and
oxygen on the functional groups [9–11]. Therefore, the amount of Dy

5996
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

ions adsorbed at equilibrium on the MPS (3, 5, 12, and 22 nm) samples via soft templating method using organic surfactants, and the pore
functionalized with N6-APTES and diglycolic acids (−2NH-2COOH) diameter was controlled by an organic reagent. The surface of MPS was
was 5.7, 8.7, 26.6, and 32.9 μg/mg silica, respectively. From these re- functionalized using a silane coupling reagent and diglycolic acid. To
sults, it can be observed that Dy adsorption increased as the pore size determine the mesoporous character of the materials, FE-SEM, TEM,
increased. In contrast, for the MPS (3, 5, 12, and 22 nm) samples and N2 physisorption were performed. The presence of the amino and
functionalized with APTES and diglycolic acids (−NH-COOH), the carboxyl groups on the surface were confirmed via TG-DTA, FTIR, and
amount of Dy adsorbed was 2.6, 4.8, 6.4, and 7.0 μg/mg silica, re- zeta-potential. Among two types of carboxyl functionalization of MPS
spectively. It is expected that for MPS samples having higher pore size with pore size (3, 5, 12 and 22 nm), the highly Dy ion adsorption was
and higher amount of organic functional groups (i.e., MPS (12 nm)- shown for MPS (12 and 22 nm)-2NH-2COOH with pore sizes of 12 and
2NH-2COOH and MPS (22 nm)-2NH-2COOH), the amount of adsorbed 22 nm. This was due to the effect of pore size on MPS, as shown by N2
Dy greatly increased up to 26 and 32 μg/mg, respectively. According to physisorption results in Fig. 3. The pore diameters of MPS (12 and
a previous report, materials such as silica gel have a Dy metal ion ad- 22 nm) provide sufficient space for Dy adsorption after functionaliza-
sorption capacity of 0.146 mmoL/g (24 μg/mg) [10]. Our results illu- tion with carboxyl groups; however, smaller pore sizes of 3 and 5 nm
strated that MPS (22 nm)-2NH-2COOH, the adsorption capacity in- were filled with the amino and carboxyl functional groups that hin-
creased more than 130% in comparison with that reported for silica gel. dered the adsorption of metal ions. These results indicated that the
The MPS-2NH-2COOH sample with smaller pore sizes (3 and 5 nm) investigation of the pore size along with a high degree of functionali-
exhibited the lowest adsorption capacity (5.7 and 8.7 μg/mg, respec- zation could improve the adsorption of Dy ions.
tively). On the contrary, MPS functionalized with the amino and car-
boxyl groups (−2NH-2COOH) with pore sizes of 12 and 22 nm had Acknowledgement
higher Dy adsorption capacity. These results indicate that the amount of
the functional groups on MPS is considerably important for ensuring the This work is partly supported by Grant-in-Aid for Scientific Research
high adsorption capacity of these materials toward Dy ions. Based on (C) No. 15K06474 from the Japan Society for the Promotion of Science
our experimental findings, a schematic of the relation between Dy ad- (JSPS).
sorption and surface functionalization of MPS samples is shown in
Fig. 8. It can be seen that the pores with a diameter of 3 nm were References
collapsed by the organic groups attached on their internal surface, and
there is no space in the pores. Therefore, it was difficult to retain the [1] J. Aguado, J.M. Arsuaga, A. Arencibia, M. Lindo, V. Gascon, Aqueous heavy metals
functionalization groups. In conclusion, it can be affirmed that the pore removal by adsorption on amine-functionalized mesoporous silica, J. Hazard.
Mater. 163 (2009) 213–221.
size and organic functional groups on MPS greatly affect the adsorption [2] A.M. Liu, K. Hidajat, S. Kawi, Y. Zhao, A new class of hybrid mesoporous materials
of Dy ions. with functionalized organic monolayers for selective adsorption of heavy metal
ions, Chem. Commun. (2000) 1145–1146.
[3] A. Sayari, S. Hamoudi, Y. Yang, Applications of pore-expanded mesoporous silica. 1.
3.4. Adsorption isotherm of Dy ions on MPS samples Removal of heavy metals cations and organic pollutants from wastewater, Chem.
Mater. 17 (2005) 212–216.
The results for the Dy adsorption capacity of the functionalized MPS [4] P.K. Jal, S. Patel, B.K. Mishra, Chemical modification of silica surface by im-
mobilization of functional groups for extractive concentration of metal ions, Talanta
samples (12 and 22 nm) are summarized in Table 2. The adsorption 62 (2004) 1005–1028.
models were evaluated from the Langmuir and Freundlich plots. The [5] T.A. Saleh, A. Sar, M. Tuzen, Optimization of parameters with experimental design
Langmuir model showed better fit than the Freundlich model with re- for the adsorption of mercury using polyethylenimine modified-activated carbon, J.
Environ. Chem. Eng. 5 (2017) 1079–1088.
gard to the R2 value. As summarized in Table 2, the R2 values for the
[6] T.A. Saleh, M. Tuzen, A. Sarı, Magnetic activated carbon loaded with tungsten oxide
Langmuir plots corresponding to MPS (12 nm)-2NH-2COOH and MPS nanoparticles for aluminum removal from waters, J. Environ. Chem. Eng. 5 (2017)
(22 nm)-2NH-2COOH were the same, i.e., 0.99, showing a good fit. The 2853–2860.
R2 values for Freundlich plots corresponding MPS (12 nm) -2NH- [7] T.A. Saleh, Simultaneous adsorptive desulfurization of diesel fuel over bimetallic
nanoparticles loaded on activated carbon, J. Clean. Prod. 172 (2018) 2123–2132.
2COOH and MPS (22 nm)-2NH-2COOH were lower (0.95 and 0.94, [8] T.A. Saleh, Mercury sorption by silica/carbon nanotube and silica/activated carbon:
respectively). From these results, it can be stated that Dy ions are ad- acomparison study, J. Water Supply Res. Technol. 64 (2015) 892–903.
sorbed on these functionalized MPS materials as a monolayer. The ex- [9] N. Zhang, B. Hu, C. Huang, A new ion-imprinted silica gel sorbent for on-line se-
lective solid-phase extraction of dysprosium (Ⅲ) with detection by inductively
perimental maximum Dy adsorption capacity for MPS (12 nm)-2NH- coupled plasma-atomic emission spectrometry, Anal. Chim. Acta 597 (2007) 12–18.
2COOH and MPS (22 nm)-2NH-2COOH (Qm) calculated using the [10] T. Ogata, H. Narita, M. Tanaka, Adsorption behavior of rare earth elements on silica
Langmuir model was 39.1 and 44.8 (μg/mg), respectively. Thus, these gel modified with diglycol amic acid, Hydrometallurgy 152 (2015) 179–182.
[11] X. Zheng, E. Liu, F. Zhang, Y. Yan, J. Pan, Efficient adsorption and separation of
results also confirm that higher pore sizes are more favorable for Dy dysprosium from NdFeB magnets in an acidic system by ion imprinted mesoporous
adsorption capacity. silica sealed in a dialysis bag, Green Chem. 18 (2016) 5031–5040.
[12] M. Anbia, K. Kargosha, S. Khoshbooei, Heavy metal ions removal from aqueous
media by modified magnetic mesoporous silica MCM-48, Chem. Eng. Res. Des. 93
3.5. Cycling test of adsorption on MPS
(2015) 779–788.
[13] C. Liu, Y. Guo, Q. Hong, C. Rao, H. Zhang, Y. Dong, L. Huang, X. Lu, N. Bao, Bovine
Cycling efficiency is one of the most important factors when col- serum albumin adsorption in mesoporous titanium dioxide: pore size and pore
chemistry effect, Langmuir 32 (2016) 3995–4003.
lection materials are used as resource adsorbents. In the cycling test, an
[14] J.Y. Ying, C.P. Mehnert, M.S. Wong, Synthesis and applications of supramolecular-
experiment was performed using MPS (22 nm)-2NH-2COOH. The ad- templated mesoporous materials, Angew. Chem. Int. Ed. 38 (1999) 56–77.
sorption process was performed according to the procedure described in [15] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky,
Section 2.7. The desorption process was performed using 1 M HNO3. Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 ang-
strom pores, Science 279 (1998) 548–552.
The cycling test was performed five times, the results of which are [16] F. Gao, C. Lian, L. Zhou, H. Liu, J. Hu, Phase separation of mixed micelles and
shown in Fig. 9. It can be observed that the adsorption capacity did not synthesis of hierarchical porous materials, Langmuir 30 (2014) 11284–11291.
change after five cycles. [17] B.G. Trewyn, I. Slowing, S. Giri, H.T. Chen, V.S.Y. Lin, Synthesis and functionali-
zation of a mesoporous silica nanoparticle based on the sol-gel process and appli-
cations in controlled release, Acc. Chem. Res. 40 (2007) 846–853.
4. Conclusion [18] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartull, J.S. Beck, Ordered meso-
porous molecular sieves synthesized by a liquid-crystal template mechanism,
Nature 359 (1992) 710–712.
Herein, we reported that the rare-earth element Dy can be suc- [19] H. Choi, J.J. Kim, Y.H. Mo, B.M. Reddy, S.E. Park, Novelty of dynamic process in
cessfully adsorbed on the amino- and carboxyl-functionalized MPS the synthesis of biocompatible silica nanotube by biomimetic glycyldodecylamide
materials with different pore sizes. These materials were synthesized as a soft template, Langmuir 33 (2017) 10707–10714.

5997
T. Kaneko et al. Journal of Environmental Chemical Engineering 6 (2018) 5990–5998

[20] K. Nakanishi, M. Tomita, K. Kato, Synthesis of amino-functionalized mesoporous adsorption by peat in single and binary component system, J. Colloid Int. Sci. 280
silica sheets and their application for metal ion capture, J. Asian Ceram. Soc. 3 (2004) 322–333.
(2015) 70–76. [27] A. Kapoor, J.A. Ritter, R.T. Yang, An extended Langmuir model for adsorption of
[21] T. Orita, M. Tomita, M. Harada, K. Kato, Binding activity of avidin to the biotin gas mixtures on heterogeneous surfaces, Langmuir 6 (1990) 660–664.
within mesoporous silica materials for bioanalytical applications, Anal. Biochem. [28] M.D. Levan, T. Vermeulen, Binary Langmuir and Freundlich isotherms for ideal
425 (2012) 1–9. adsorbed solutions, J. Phys. Chem. 85 (1981) 3247–3250.
[22] J.L. Paris, M. Colilla, L.L. Barba, M. Manzano, M.V. Regi, Tuning mesoporous silica [29] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers, J.
dissolution in physiological environments: a review, J. Mater. Sci. 52 (2017) Am. Chem. Soc. 60 (1938) 309–319.
8761–8771. [30] A. Suwalski, H. Dabboue, A. Delalande, S.F. Bensamoun, F. Canon, P. Midoux,
[23] K. Nakanishi, M. Tomita, K. Kato, Improvement in the catalytic activity of cyto- G. Saillant, D. Klatzmann, J. PSalvetat, C. Pichon, Accelerated achilles tendon
chrome c by immobilization on a novel mesoporous silica sheet, RSC Adv. 4 (2014) healing by PDGF gene delivery with mesoporous silica nanoparticles, Biomaterials
4732–4735. 31 (2010) 5237–5245.
[24] Y. An, M. Chen, Q. Xue, W. Liu, Preparation and self-assembly of carboxylic acid- [31] A.S.M. Chong, X.S. Zhao, Functionalization of SBA-15 with APTES and character-
functionalized silica, J. Colloid Interface Sci. 311 (2007) 507–513. ization of functionalized materials, J. Phys. Chem. B 107 (2003) 12650–12657.
[25] R. Hikosaka, F. Nagata, M. Tomita, K. Kato, Adsorption and desorption character- [32] T.A. Saleh, Isotherm, kinetic, and thermodynamic studies on Hg(II) adsorption from
istics of DNA onto the surface of amino functional mesoporous silica with various aqueous solution by silica- multiwall carbon nanotubes, J. Water Supply Res.
particle morphologies, Colloids Surf. B Biointerfaces 140 (2016) 262–268. Technol. 21 (2015) 16721–16731.
[26] S.J. Allen, G. Mckay, J.F. Porter, Adsorption isotherm models for basic dye

5998

Potrebbero piacerti anche