You are on page 1of 18

20230831101110j5rD

Aug 31, 2023


4183 words / 23634 characters

20230831101110j5rD

Sources Overview

26%
OVERALL SIMILARITY

Jun Su, Ying Jia, Menglin Shi, Keke Shen, Jiqing Zhang. "Highly e cient unsymmetrical dimethylhydrazine removal from wastewater using MIL-53(Al)-derived ca…
1
CROSSREF
9%

Zikang Xiong, Huaili Zheng, Yadan Hu, Xuebin Hu, Wei Ding, Jiangya Ma, Yisen Li. "Selective adsorption of Congo red and Cu(II) from complex wastewater by cor…
2
CROSSREF
3%

Chen Wang, Chao Xiong, Xusheng Zhang, Yali He, Jianhua Xu, Yurong Zhao, Shixing Wang, Jianzhong Zheng. "External optimization of Zr-MOF with mercapto…
3
CROSSREF
<1%

Siqi Li, Xiaodan Zhang, Yuming Huang. "Zeolitic imidazolate framework-8 derived nanoporous carbon as an effective and recyclable adsorbent for removal of c…
4
CROSSREF
<1%

Yihan Wu, Biyun Li, Xiangxue Wang, Shujun Yu, Hongwei Pang, Yue Liu, Xiaoyan Liu, Xiangke Wang. "Magnetic metal-organic frameworks (Fe3O4@ZIF-8) com…
5
CROSSREF
<1%

iwaponline.com
6
INTERNET
<1%

prr.hec.gov.pk
7
INTERNET
<1%

pubs.rsc.org
8
INTERNET
<1%

Huaiyuan Niu, Xueting Li, Jishan Li. "Dithiocarbamate Modi cation of Activated Carbon for the E cient Removal of Pb(II), Cd(II), and Cu(II) from Water", New …
9
CROSSREF
<1%

coek.info
10
INTERNET
<1%

George Z. Kyzas, Panoraia I. Siafaka, Dimitra A. Lambropoulou, Nikolaos K. Lazaridis, Dimitrios N. Bikiaris. "Poly(itaconic acid)-Grafted Chitosan Adsorbents …
11
CROSSREF
<1%
Yuhua Niu, Xingxing Han, Liangxian Huang, Jie Song. "Methylene Blue and Lead(II) Removal via Degradable Interpenetrating Network Hydrogels", Journal of C…
12
CROSSREF
<1%

link.springer.com
13
INTERNET
<1%

Neda Asasian, Tahereh Kaghazchi. "Optimization of Activated Carbon Sulfurization to Reach Adsorbent with the Highest Capacity for Mercury Adsorption", Se…
14
CROSSREF
<1%
Yanhui Zhan, Mingyue Chang, Jianwei Lin. "Suppression of phosphorus release from sediment using lanthanum carbonate as amendment", Environmental Sci…
15
CROSSREF
<1%
Yizhen Cheng, Binyuan Wang, Jimin Shen, Pengwei Yan et al. "Preparation of novel N-doped biochar and its high adsorption capacity for atrazine based on π–…
16
CROSSREF
<1%
Yanshan Yin, Jie Yin, Wei Zhang, Hong Tian, Zhangmao Hu, Min Ruan, Ziyou Song, Liang Liu. "Effect of Char Structure Evolution During Pyrolysis on Combustio…
17
CROSSREF
<1%
"Green Adsorbents to Remove Metals, Dyes and Boron from Polluted Water", Springer Science and Business Media LLC, 2021
18
CROSSREF
<1%
core.ac.uk
19
INTERNET
<1%
www.science.gov
20
INTERNET
<1%
Wensheng Linghu, Hai Yang, Yanxia Sun, Guodong Sheng, Yuying Huang. "One-Pot Synthesis of LDH/GO Composites as Highly Effective Adsorbents for Deco…
21
CROSSREF
<1%
Jiahui Zhu, Qi Liu, Jingyuan Liu, Rongrong Chen, Hongsen Zhang, Jing Yu, Milin Zhang, Rumin Li, Jun Wang. "Novel Ion-Imprinted Carbon Material Induced by …
22
CROSSREF
<1%
Mohamed A. Farghali, Ramadan M. El Bahnasawy, Taher A. Salaheldin. "A magnetic graphene nanocomposite for e cient removal of methylene blue from wa…
23
CROSSREF
<1%
Rong-Zhong Wang, Dan-Lian Huang, Yun-Guo Liu, Chen Zhang et al. "Synergistic removal of copper and tetracycline from aqueous solution by steam-activate…
24
CROSSREF
<1%
Yilin Chen, Yong Qin, Chongtao Wei, Lili Huang, Qingmin Shi, Caifang Wu, Xiaoyang Zhang. "Porosity changes in progressively pulverized anthracite subsample…
25
CROSSREF
<1%
Zhang, K.j.. "Granular activated carbon (GAC) adsorption of two algal odorants, dimethyl trisul de and @b-cyclocitral", Desalination, 20110131
26
CROSSREF
<1%
cancer-nano.biomedcentral.com
27
INTERNET
<1%
research.abo.
28
INTERNET
<1%
www.mdpi.com
29
INTERNET
<1%
C. Namasivayam, D. Sangeetha. " Removal of chromium(VI) by ZnCl activated coir pith carbon ", Toxicological & Environmental Chemistry, 2006
30
CROSSREF
<1%
H. Faghihian, H. Nourmoradi, M. Shokouhi. "Removal of copper (II) and nickel (II) from aqueous media using silica aerogel modi ed with amino propyl triethoxy…
31
CROSSREF
<1%
irgu.unigoa.ac.in
32
INTERNET
<1%
Preprint source
33 www.researchsquare.com <1%
INTERNET

www.sciencegate.app
34
INTERNET
<1%
Abdelmajid Regti, My Rachid Laamari, Salah-Eddine Stiriba, Mohammadine El Haddad. " Potential use of activated carbon derived from species under alkaline …
35
CROSSREF
<1%
Eren, Z.. "Equilibrium and kinetic mechanism for Reactive Black 5 sorption onto high lime Soma y ash", Journal of Hazardous Materials, 20070508
36
CROSSREF
<1%
M. Shabani, F. Azizinezhad. "Adsorption of Acid Blue 260 from Aqueous Solutions onto Multi-Wall Carbon Nanotube: Determination of Equilibrium, Thermodyn…
37
CROSSREF
<1%
hjxy.jnu.edu.cn
38
INTERNET
<1%
mdpi-res.com
39
INTERNET
<1%
revues.imist.ma
40
INTERNET
<1%
Wenjihao Hu, Xia Wu, Feipeng Jiao, Weijie Yang, Yangmeihui Zhou. " Preparation and characterization of magnetic Fe O @sulfonated -cyclodextrin intercalated…
41
CROSSREF
<1%
rd.springer.com
42
INTERNET
<1%
repository.ntu.edu.sg
43
INTERNET
<1%
yp.chem.tue.nl
44
INTERNET
<1%
Wen-Jun Luo, Qiang Gao, Xiao-Liang Wu, Cheng-Gang Zhou. "Removal of Cationic Dye (Methylene Blue) from Aqueous Solution by Humic Acid-Modi ed Expan…
45
CROSSREF
<1%
Songlin Chao, Fang Zou, Fanfan Wan, Xiaobin Dong, Yanlin Wang, Yuxuan Wang, Qingxin Guan, Guichang Wang, Wei Li. "Nitrogen-doped Carbon Derived from …
46
CROSSREF
<1%

Excluded search repositories:


None

Excluded from document:


Bibliography

Excluded sources:
None

Excluded preprints
None
32
1 Results and discussion

2 3.1 Materials characterizations

3 3.1.1 Morphology characteristics

20
4 Scanning electron microscope (SEM) and transmission electron microscopy (TEM) were utilized

4
5 to analyze the morphology, microstructure, and elemental distribution of ZIF-8 and FCZ@C-600. The

6 synthesized ZIF-8 exhibits a rhombic dodecahedral shape (Fig. 1a, b, and e), like the shape described

4
7 in previous reports [1,2]. The particle size of ZIF-8 is uniform, with an average diameter of ~600 nm.

8 SEM (Fig. 1c and d) and TEM (Fig. 1f and g) show that the Fe3O4 nanoparticles uniformly coat the

2
9 surface of ZIF-8 and retain their magnetic properties. The boundary between the amorphous CS shell

10 and ZIF-8 is indistinct, possibly due to the small mass difference between the two components [3].

11 The high-resolution TEM image (Fig. 1h) shows that the interplanar lattice spacing is ~0.47 nm, which

12 can be assigned to the spacing of the (111) crystal plane of Fe3O4 [4]. The inset in Fig. 1h shows a

13 selected area electron diffraction (SAED) pattern of FCZ@C-600, which shows a series of diffraction

14 rings corresponding to the crystal planes of the polycrystalline Fe3O4 phase, consistent with the XRD

15 analysis. The presence of C, N, Zn, Fe, and O in FCZ@C-600 was confirmed by EDS analysis (Fig.

16 1i). C, N, and Zn originate from the imidazole rings and metal clusters of ZIF-8, while Fe and O are

17 mainly from Fe3O4. This indicates that Fe3O4 nanoparticles are scatteredly incorporated in ZIF-8.
18
13 46
19 Fig. 1. SEM images of (a, b) ZIF-8 and (c, d) FCZ@C-600; TEM image of (e) ZIF-8 and (f−h) FCZ@C-600 under different
16
20 magnifications (the inset in Fig. 1Fig. 1h shows the corresponding SAED pattern of FCZ@C-600); (i) EDS elemental Formatted: Fon
21 distribution of C, N, Zn, Fe, and O in FCZ@C-600.

22 3.1.2 C rystallographic structure

16 4
23 The X-ray diffraction (XRD) patterns of synthesized Fe3O4 and ZIF-8 are consistent with the

1
1 simulated patterns (Fig. 2a). The (220), (311), (400), (422), (511) and (440) diffraction peaks at 30.13°,

2 35.46°, 43.18°, 53.57°, 57.14° and 62.72° belong to highly crystalline Fe3O4 (JCPDS No. 3−863)

3 [5,6]. After introducing CS and ZIF-8, the characteristic Fe3O4 peaks in FCZ and FCZ@C-600 remain,

4 indicating that the modification and pyrolysis do not destroy the Fe3O4 structure. Moreover, the XRD

5 patterns of ZIF-8 and FCZ show distinct peaks at 7.48°, 10.41°, 12.87°, 14.87°, 16.62°, 18.16°, 22.34°,

6 24.69°, and 26.91°, corresponding to planes (011), (002), (112), (022), (013), (222), (114), (223), and

7 (134) of the ZIF-8 crystal (CCDC 864309), respectively [7]. This confirms the accurate preparation of

43
8 ZIF-8 and the successful loading of ZIF-8 on the surface of FC. The ZIF-8 phase in FCZ crystallized

9 into the ZnO phase in FCZ@C-600 (JCPDS No. 21−1486) [8,9], suggesting the conversion of ZIF-8

10 into ZnO products after pyrolysis at 600 °C.


11
28
12 Fig. 2. (a) XRD spectra of Fe3O4, ZIF-8, FCZ, and FCZ@C-600; (b) FT-IR spectra of CS, Fe3O4, FC, ZIF-8, FCZ, and
13 FCZ@C-600; (c) Magnetization curves of FCZ and FCZ@Cs (the inset shows the dispersion and magnetic separation of
8
14 FCZ@C-600 in aqueous solution); (d) Thermogravimetric curves of ZIF-8, Fe3O4, FC, and FCZ; (e) N2
15 adsorption−desorption isotherms; and (f) pore diameter distribution profiles of FCZ and FCZ@C-600.

33
16 3.1.3 Surface functional groups

17 Fourier transform infrared spectra (FT-IR) were obtained to evaluate the functional groups (Fig.

21
18 2b). In the FT-IR spectrum of CS and FC, the peaks at 1089, 1385, and 1634 cm−1 correspond to the

8
19 C−O−C stretching, C−H bending, and C=O stretching vibrations, respectively, in CS [10]. The

20 characteristic band at 568 cm−1 can be attributed to the Fe−O stretching vibration [11], indicating the

21 formation of Fe3O4 nanoparticles. These characteristic Fe3O4 peaks are observed in the FC, FCZ, and

2
22 FCZ@C-600 spectra, indicating the existence of magnetic Fe3O4 in the adsorbents. The bands

23 observed in the 600−800, 900−1340, and 1340−1550 cm−1 regions of the ZIF-8 spectrum are

2
1 attributed to the out-of-plane vibration, in-plane vibration and stretching vibration of the imidazole

2
2 ring [12,13]. In particular, the C=N stretching mode at 1585 cm−1 and peaks associated with the

3 aromatic and aliphatic C−H stretching vibrations at 2917 and 3140 cm−1, respectively, are also

27
4 observed [14]. Moreover, the peak at 423 cm−1 is attributed to the Zn−N stretching vibration [15].

5 These characteristic peaks are observed in the FCZ spectrum and indicate successful coating of ZIF-8.

2
6 3.1.4 Magnetic hysteresis curves and thermogravimetric analysis

7 The saturation magnetization and hysteresis loops were measured with a vibrating sample

2 2
8 magnetometer to investigate the separation performance of the adsorbents. In all curves, symmetrically

9 distributed S-shaped curves pass through the origin without significant hysteresis (Fig. 2c), indicating

10 that the adsorbents are typical superparamagnetic materials [16]. The magnetic saturation values of

11 FCZ and FCZ@C-600 are ~17.6 and 114.6 emu/g, respectively. The relatively weak magnetism of

12 FCZ is mainly affected by coating with non-magnetic carbonaceous material (CS and ZIF-8), which

13 further demonstrates the successful synthesis of the composite particles. In addition, pyrolysis

14 significantly increases the magnetic saturation values of the adsorbents. This effect increases with

15 increasing pyrolysis temperature as pyrolysis decomposes the non-magnetic carbonaceous material,

16 thus increasing the mass fraction of magnetic Fe3O4. In addition, FCZ@C-600 could be dispersed

5
17 very well in aqueous solutions due to its submicron particle size but responded rapidly to an external

18 magnetic field. These results indicate that FCZ@C-600 can be easily separated, an essential property

19 of environmentally friendly adsorbents for environmental applications.

2
20 The saturation magnetization and hysteresis loops were measured with a vibrating sample

2 2
21 magnetometer to investigate the separation performance of the adsorbents. In all curves, symmetrically

3
1 distributed S-shaped curves pass through the origin without significant hysteresis (Fig. 2c), indicating

2 that the adsorbents are typical superparamagnetic materials [16]. The magnetic saturation values of

3 FCZ and FCZ@C-600 are ~17.6 and 114.6 emu/g, respectively. The relatively weak magnetism of

4 FCZ is mainly affected by coating with non-magnetic carbonaceous material (CS and ZIF-8), which

5 further proves the successful synthesis of the composite particles. Moreover, pyrolysis significantly

6 increases the magnetic saturation values of the adsorbents. This effect increases with increasing

7 pyrolysis temperature as pyrolysis decomposes the non-magnetic carbonaceous material, thus

8 increasing the percentage mass of magnetic Fe3O4. In addition, FCZ@C-600 could be dispersed very

5
9 well in aqueous solutions due to its submicron particle size but responded rapidly to an external

10 magnetic field. These results indicate that FCZ@C-600 can be easily separated, an essential quality

11 property of environmentally friendly adsorbents for environmental applications.

30
12 3.1.5 Specific surface area and pore diameter distribution

2
13 Geometric characteristics were measured in the N2 adsorption−desorption analysis, including

14 specific surface area (SBET), pore volume, and average pore diameter of FCZ and FCZ@C-600 (Figs.

41
15 2e and f). FCZ and FCZ@C-600 show similar isotherms of typical type IV with obvious hysteresis

16 loops of H3 type at P/P0 of 0.4−0.99, indicating the existence of micropores and mesoporous pores

17 [18]. The SBET of FCZ@C-600 (85.69 m2/g) is much smaller than that of FCZ (327.57 m2/g), which

2
18 is mainly due to the heavier and nonporous magnetic Fe3O4 after pyrolysis at 600 °C. Moreover, the

1
19 pore volume and diameter increase after pyrolysis due to the decomposition of the ligands (Table 1),

20 resulting in a hierarchical porous structure that provides a larger contact area and facilitates the mass

21 transfer of the desired contaminants [12].

4
1 Table 1. Pore structure parameters of FCZ and FCZ@C-600.
19
Adsorbents SBET (m2/g)a Pore volume (cm3/g)b Average pore diameter (nm)c
FCZ 327.57 0.2062 10.32
FCZ@C-600 85.69 0.2804 16.68
42
2 SBET (specific surface area) was analyzed with the BET method.
a

25
3 bThe pore volume was calculated by using adsorption data and the BJH method.
4 The average pore diameter was calculated using adsorption data and the BJH method.
c

2
5 3.2 Effect of system parameters on adsorption

6 3.2.1 Effect of different adsorbents

7 Comparative radar plots are often required to evaluate adsorption performance at different

35
8 pyrolysis temperatures. Fig. 3a shows the effects of pyrolysis temperatures on the adsorption capacity

17
9 (mg/g) of FC, FCZ, and ZIF-8. The adsorption capacity of ZIF-8@Cs increases with increasing

10 pyrolysis temperature due to the decomposition of the ligand, resulting in additional pores [19].

17
11 Interestingly, the adsorption capacity of FC @Cs and FCZ@Cs first decreases and then increases with

12 increasing pyrolysis temperature, reaching a maximum of 600 °C. The adsorption capacity increases

13 significantly from 200 to 600 °C, which could be attributed to the larger porosity and pore diameter of

14 FC @Cs and FCZ@Cs after pyrolysis. However, above 600 °C, the adsorption capacity starts to

15 decrease as the carbon layer derived from CS /ZIF-8– with high adsorption performance is gradually

16 lost with increasing pyrolysis temperature, while the relative mass of magnetic Fe3O4 with weak

2
17 adsorption performance increases. These results show that the adsorption performance of FC @Cs and

2
18 FCZ@Cs is mainly based on CS /ZIF-8–derived porous carbons and prove that the successful

2
19 attachment of CS and ZIF-8 increases the effective adsorption sites on the adsorbents.

1
20 The adsorption capacity and removal efficiency data of FCZ, FCZ@Cs, AC, and GO are shown

1 7
21 in Fig. 3b. FCZ@C-600 shows the highest adsorption capacity (44.51 mg/g), which is 1.72 times

5
1 higher than that of GO and 1.58 times higher than that of AC. Characterizations of the physical and

2 chemical properties of AC and GO are shown in Figs. S2 and S3, respectively. In addition, ZIF-8@C-

3 600 and FCZ@C-600 have similar adsorption capacities and are both magnetically recyclable.

4 Therefore, considering the practical application and economic cost, FCZ@C-600 was selected to study

5 the effects of system parameters on adsorption. Indeed, FCZ@C-600 could be a competitive adsorbent

6 for removing UDMH from wastewater.


7
8 Fig. 3. (a) Comparative radar plots of the adsorption capacity (mg/g) of FC, FCZ, and ZIF-8 at different pyrolysis
1
9 temperatures; (b) Comparison of the adsorption capacity and removal efficiency of different adsorbents. Detailed
9
10 parameters: [UDMH] = 50 mg/L, T = 298 K, m/V = 0.8 g/L, t = 24 h, initial pH = 8.4.

1
11 3.2.2 Effect of temperature, initial concentration, and dosage

12 The effect of temperature on the adsorption affinity of FCZ@C-600 and ZIF-8@C-600 for

1
13 UDMH was also investigated (Fig. 4). From 288 to 318 K, the improved contact probability between

14 UDMH and the adsorption sites led to a gradual increase in both adsorption capacity and removal

1
15 efficiency [20]. Thus, the increase in temperature promotes the adsorption of UDMH, which is

18
16 consistent with the thermodynamic results (ΔH0 > 0) and shows that the adsorption process is

17 endothermic.

12
18 At a fixed dosage (0.8 g/L) and increasing initial UDMH concentration (25(500 mg/L), the

19 adsorption capacity of FCZ@C-600 and ZIF-8@C-600 improves significantly and eventually

20 approaches equilibrium. At the same time, the removal efficiency decreases gradually (Figs. 4a and b).

21 The adsorption properties indicate high concentrations favor molecule collisions and promote

22 adsorption [20,21].

23 To select the optimal dosage for the removal of UDMH, different amounts of FCZ@C-600 and

6
36
1 ZIF-8@C-600 were studied at a fixed concentration (50 mg/L). The adsorption capacity gradually

9
2 decreases with increasing dosage (0.2(1.4 g/L), while the removal efficiency gradually increases (Fig.

1
3 4c and d). A reasonable explanation is that the effective adsorption sites increase with increasing

4 dosage [22]. The change of FCZ@C-600 dosage from 0.2 to 0.8 g/L at 298 K increases the UDMH

5 removal efficiency from 45.91% to 71.78%. Further, increasing the dosage of FCZ@C-600 and ZIF-

6 8@C-600 has no significant effect on the removal efficiency. Therefore, 0.8 g/L is considered the

7 optimum dosage for removal efficiency and economy.


8
23
9 Fig. 4. (a) Effect of temperature and initial concentration on the adsorption capacity of FCZ@C-600 and ZIF-8@C-600;
37 1
10 (b) Effect of temperature and initial concentration on the removal efficiency of FCZ@C-600 and ZIF-8@C-600; (c) Effect
1
11 of temperature and dosage on the adsorption capacity of FCZ@C-600 and ZIF-8@C-600; (d) Effect of temperature and
1
12 dosage on the removal efficiency of FCZ@C-600 and ZIF-8@C-600. Detailed parameters: T = 288, 298, 308, and 318 K,
13 t = 16 h, initial pH = 8.4.

26
14 3.2.3 Effect of I nitial pH

15 The initial pH can affect the adsorbents' surface charge and the contaminants' molecular

1
16 morphology [23]. The adsorption capacity of FCZ@C-600 is affected by the initial pH (Fig. 5a),

17 indicating the presence of electrostatic interactions. It is noteworthy that the pKa of UDMH is 8.38

1
18 [24]; the percentage of the deprotonated form increases sharply from 0% to 99.5% when the pH

1
19 increases from 6 to 11 (Fig. 5b). Moreover, the point of zero charge (pHPZC) for FCZ@C-600 is 5.6,

24
20 so when the pH < is 5.6, the UDMH molecules are completely protonated and the surface of the

21 adsorbent is positively charged. Accordingly, the electrostatic repulsion between protonated UDMH

1
22 and positively charged FCZ@C-600 results in an extremely low adsorption capacity [25]. The

23 electrostatic attraction between fully protonated UDMH and negatively charged FCZ@C-600 further

18
24 contributes to the increase in adsorption capacity when the pH increases to 6. When the pH is further

7
1 increased, the adsorption capacity decreases, which may be caused by the weakening of electrostatic

2 attraction due to the rapid deprotonation of UDMH and the increase in the number of competing free

3 hydroxide ions [12].


4
1 1
5 Fig. 5. (a) Effect of the initial pH on the adsorption capacity and zeta potential of FCZ@C-600; (b) Distribution of
1
6 deprotonated forms of UDMH under different pH conditions; (c) Effect of the ionic strength, (d) humic acid, (e) co-existing
9
7 ions, and (f) water qualities on the adsorption capacity of FCZ@C-600. Detailed parameters: [UDMH] = 50 mg/L, T = 298
8 K, m/V = 0.8 g/L, t = 16 h.

1
9 3.2.4 Effect of ionic strength, HA, co-existing ions, and water qualities

1
10 Since water contains many impurities, natural organic matter (NOM) and interfering ions were

1
11 treated as additives to study their effects on the adsorption of UDMH. However, ionic strength plays a

12 limited role in UDMH adsorption (Fig. 5c), suggesting that electrostatic shielding has a limited effect

13 on adsorption capacity [26]. The adsorption capacity increases with increasing HA concentration up to

4
14 40 mg/L and remains relatively stable at high HA concentrations (Fig. 5d). This is due to HA, which

4 4
15 can be adsorbed by carbon materials and provides additional active sites by forming hydrogen bonds

16 between the abundant hydroxyl groups and amino groups of UDMH [27]. However, due to HA's low

17 adsorption capacity, further increasing HA concentrations does not positively affect the adsorption

18 performance. Fig. 5e shows that different coexisting ions inhibit the adsorption of UDMH. The

19 inhibition increases with ion concentration, suggesting that these ions compete with UDMH for

20 adsorption sites [11]. Further adsorption experiments were conducted under simulated natural

21 conditions of real UDMH wastewater. The adsorption capacity of FCZ@C-600 does not vary

22 significantly with different water qualities and can be ranked as follows: Ultrapure water > Tap water >

23 Synthetic surface water > Lake water (Fig. 5f). Therefore, FCZ@C-600 is an effective agent for

8
1 treating natural water containing interfering ions and NOM with environmental stability.

29
2 3.3 Adsorption performance

3 3.3.1 A dsorption kinetics

14
4 The effect of contact time on the adsorption capacity of FCZ@C-600 and ZIF-8@C-600 was

10
5 determined by varying the adsorption time (Fig. 6a). At the initial UDMH concentrations of 50 and

6 100 mg/L, the adsorption processes of FCZ@C-600 and ZIF-8@C-600 increase rapidly within 9 hours.

7 Subsequently, the adsorption process slows down and approaches equilibrium after 16 hours. Initially,

14
8 the adsorption capacity of FCZ@C-600 is higher than that of ZIF-8@C-600, indicating that FCZ@C-

9 600 has a faster adsorption rate, which is consistent with the kinetic rate constants (Table 2). According

10 to these, the adsorption capacity of ZIF-8@C-600 at concentrations of 50 and 100 mg/L exceeds that

11 of FCZ@C-600 after ~9 and 12 hours, respectively, which may be due to the higher equilibrium

12 adsorption capacity of ZIF-8@C-600.

7
13 To gain insight into the adsorption mechanisms, the pseudo-first-order (PFO) and pseudo-second-

3
14 order (PSO) kinetics models were used to fit the kinetics data. The nonlinear fitting curves of the PFO

13
15 and PSO models are shown in Figs. 6b and c and the corresponding kinetics parameters are listed in

16 Table 2. The PFO model assumes that the diffusion step controls the adsorption process [28]. The PSO

17 model also assumes that the adsorption rate is controlled by chemical interactions, including electron

18 pair sharing [28]. The R2 value obtained with the PSO model is much higher than that derived from

1
19 the PFO model. Therefore, the PSO model better describes the adsorption process of UDMH

1
20 adsorption on FCZ@C-600 and ZIF-8@C-600, indicating that the rate-limiting step is chemisorption

21 involving electron sharing. The velocity parameters (k1 and k2) of FCZ@C-600 are higher than those

9
1
1 of ZIF-8@C-600, indicating that FCZ@C-600 has a faster adsorption rate.

38
2 The intraparticle diffusion model, which can explore the rate-limiting step in the adsorption

12 1
3 process, was also used to fit the kinetics data. The results show that the rate-limiting steps can be

4 controlled by multistep processes (Fig. 6d), such as surface adsorption and intraparticle diffusion

22
5 because neither the second nor the third step exceeds the original point [29]. Moreover, the constant C

15
6 value (reflecting the thickness of the boundary layer) of stage 2 is higher than that of stage 1 in each

7 adsorption process (Table S3), indicating that intraparticle diffusion plays an important role [7].

8 According to the results of BET analysis, the FCZ@C-600 magnetic adsorbent has a larger pore

9 volume and pore diameter. The presence of these pores affects the mass transfer phenomena and can

10 increase the mass transfer coefficient [30].


11
31
12 Fig. 6. (a) Effect of the contact time on the adsorption capacity of FCZ@C-600 and ZIF-8@C-600; PFO and PSO models
13 of (b) FCZ@C-600 and (c) ZIF-8@C-600; (d) Intra-particle diffusion model of FCZ@C-600 and ZIF-8@C-600. Detailed
1
14 parameters: [UDMH] = 50 and 100 mg/L, T = 298 K, m/V = 0.8 g/L, initial pH = 8.4.
15 Table 2. Kinetics parameters of UDMH adsorption onto FCZ@C-600.

10 PFO PSO
Adsorbents Initial concentration (mg/L)
qe (mg/g) k1 (h )
1
R2
qe (mg/g) k2 (g/(mg h)) R2
FCZ@C-600 50 44.32 0.4542 0.985 50.01 0.01169 0.997
100 73.92 0.2985 0.967 85.58 0.00434 0.984
ZIF-8@C-600 50 50.07 0.2515 0.973 59.07 0.00504 0.987
100 87.05 0.1371 0.945 109.92 0.00126 0.957

16 3.3.2 Adsorption isotherms

6
17 The intraparticle-intraparticle diffusion model, which can explore the rate-limiting step in the

22 12
18 adsorption process, was also used to fit the kinetics data. Results The results show that the rate-limiting

1
19 steps can be controlled by multistep processes (Fig. 6d), such as surface adsorption and intraparticle-

20 intraparticle diffusion because neither the second nor the third step exceeds the original point [29].

10
15
1 Moreover, the constant C value (reflecting the thickness of the boundary layer) of stage 2 is higher

2 than that of stage 1 in each adsorption process (Table S3), indicating that intraparticle-intraparticle

3 diffusion plays an important role [7]. According to the BET analysis results, the magnetic adsorbent

4 FCZ@C-600 has a larger pore volume and pore diameter. The presence of these pores affects the mass

5 transfer phenomena and can increase the mass transfer coefficient [30].

34
6 According to the Langmuir model, the maximum adsorption capacities (qm) of UDMH

1
7 adsorption on FCZ@C-600 and ZIF-8@C-600 are 185.70 and 213.18 mg/g at 298 K, respectively.

1
8 Table S4 compares the adsorption of UDMH on FCZ@C-600 and ZIF-8@C-600 with other reported

2
9 adsorbents. The results show that the adsorption performance of FCZ@C-600 exceeds that of most

10 adsorbents reported in the literature. Although the sq.m. value of FCZ@C-600 (185.70 mg/g) is

11 slightly lower than that of C@Al2O3-1000 (275.93 mg/g) and ZIF-8@C-600 (213.18 mg/g), effective

12 magnetic separation properties would result in an additional advantage, thus reducing cost, increasing

13 recyclability, and preventing potential secondary contamination. Thus, FCZ@C-600 could be a

14 promising candidate for UDMH removal.


15
1 1
16 Fig. 7. (a) Adsorption isotherms data of FCZ@C-600 and ZIF-8@C-600; Nonlinear fitting curves with Langmuir,
17 Freundlich, and Sips models of (b) FCZ@C-600 and (c) ZIF-8@C-600; (d) Dimensional separation factor (RL) of UDMH
18 by FCZ@C-600 and ZIF-8@C-600. Detailed parameters: [UDMH] = 25 500 mg/L, T = 288, 298, 308, and 318 K, m/V =
1
19 0.8 g/L, t = 16 h, initial pH = 8.4.
20 Table 3. Isotherms parameters of UDMH adsorption onto FCZ@C-600.
Langmuir Freundlich Sips
1
A dsorbents T (K) qm KL KF qs KS
RL R2 1/n R2 m R2
(mg/g) (L/mg) (mg/g) (mg/g) (L/mg)
FCZ@C-600 288 166.93 0.0175 0.103 0.696 0.988 16.12 0.382 0.968 198.92 0.0107 0.767 0.984
298 185.70 0.0207 0.088 0.659 0.989 20.39 0.365 0.935 193.82 0.0185 0.916 0.986
308 242.88 0.0243 0.076 0.622 0.983 28.85 0.360 0.932 244.14 0.0240 0.988 0.979
318 258.13 0.0302 0.062 0.570 0.987 35.23 0.342 0.917 256.89 0.0305 1.013 0.984

11
ZIF-8@C-600 288 139.92 0.0169 0.106 0.703 0.979 14.247 0.370 0.949 158.69 0.0121 0.806 0.980
298 213.18 0.0215 0.085 0.650 0.970 24.184 0.363 0.935 240.54 0.0157 0.800 0.970
308 249.22 0.0327 0.058 0.550 0.988 35.89 0.336 0.962 297.28 0.0195 0.718 0.983
318 262.21 0.0421 0.045 0.487 0.981 47.45 0.301 0.963 327.12 0.0213 0.648 0.978

1 3.3.3 A dsorption thermodynamics

2 Applying adsorption thermodynamics is a crucial basis for determining the adsorption behavior

3 characteristics of FCZ@C-600 and ZIF-8@C-600. Thermodynamic parameters, including the standard

11
4 change in Gibbs free energy [ΔG0, kJ/mol), the standard change in enthalpy (ΔH0, kJ/mol), and the

5 standard change in entropy (ΔS0, J/(mol ( K)], were investigated to interpret the adsorption

1
6 mechanisms further. The Van’ t Hoff equation (ln Ke0 versus 1/T) is shown in Fig. S4.

40
7 Negative ΔG0 values indicate that adsorption is favorable and spontaneous at all temperatures

1
8 studied, with a high preference for UDMH (Table 4). The ΔG0 values of ZIF-8@C-600 are more

9 negative than those of FCZ@C-600, indicating a stronger preference for adsorption of UDMH on ZIF-

10 8@C-600. This observation is consistent with the results obtained from the adsorption isotherms. The

11 endothermic character of the process is confirmed by the positive ΔH0 values, which can be attributed

1
12 to the combined effect of the desorption of water molecules and the adsorption of UDMH molecules

1
13 [33]. Moreover, the ΔS0 values are positive, meaning that the randomness at the solid–liquid interface

14 increases during adsorption [28]. Accordingly, the adsorption of UDMH on FCZ@C-600 and ZIF-

15 8@C-600 is a spontaneous, endothermic, and disordered process.


45
16 Table 4. Thermodynamic parameters for UDMH adsorption on FCZ@C-600 and ZIF-8@C-600.
6
Adsorbents T (K) K e0 ΔG 0 (kJ/mol) ΔH0 (kJ/mol) ΔS 0 (J/(mol K)) R2
FCZ@C-600 288 1049.60 16.66 13.72 105.35 0.987
298 1241.25 17.65
308 1462.35 18.66
318 1814.42 19.84

12
ZIF-8@C-600 288 1017.76 16.58 23.95 140.50 0.980
298 1293.52 17.75
308 1964.62 19.42
318 2529.37 20.72

1 3.4 A dsorption mechanisms

2 3.4.2 DFT calculation

3
3 Fig. 9d and Table S5 show the calculation results (optimal geometries, adsorption energies, and

4 bond lengths) of binding UDMH to the three selected active adsorption sites. The calculation of DFT

3
5 shows that the adsorption energies associated with the three models follow the sequence ZIF-8-

6 UDMH-1 (−0.066 eV) > ZIF-8-UDMH-2 (−0.533 eV) > ZIF-8-UDMH-3 (−0.619 eV), indicating that

7 all models are responsible for UDMH binding. Since ZIF-8-UDMH-3 has the lowest energy, UDMH

8 binding would mainly occur in this conformation [34], consistent with the results obtained for the bond

9 lengths (shorter bond lengths correspond to higher binding energy).


10
11 Fig. 89. (a) Electrostatic potential diagrams of UDMH and ZIF-8; Distribution of HOMO and LUMO orbitals of (b) UDMH
3
12 and (c) ZIF-8; (d) Optimal geometries and bond lengths of UDMH bound to three active adsorption sites on ZIF-8. Gray,
13 blue, white, and ice blue spheres represent C, N, H, and Zn atoms.

14 3.4.3 Potential mechanisms of UDMH adsorption onto FCZ@C-600

15 Research on adsorption mechanisms is critical to understanding the adsorption process. The

16 mechanisms that can be used to interpret the adsorption behavior of MOF carbons mainly include

44
17 electrostatic interactions, hydrogen bonding, π-π interactions, hydrophobic interactions, and

18 coordination of framework metals [35]. Possible mechanisms underlying the adsorption of UDMH on

19 FCZ@C-600 are discussed individually below.

20 Electrostatic interactions are associated with attraction between oppositely charged molecules or

21 adsorbent surfaces, the strength of which depends on the pH of the solution [32]. As discussed in
13
1
1 Section 3.2.3, pH significantly weakens adsorption performance at pH < 5.6, and electrostatic

2 repulsion plays a dominant role. At a pH > of 5.6, the surfaces of UDMH and FCZ@C-600 have

3 opposite charges, indicating electrostatic attraction. The maximum adsorption capacity is reached at

39
4 pH = 6 because UDMH is completely protonated, and the surface of the adsorbent has a strong negative

5 charge. At a pH of > 6, the adsorption capacity decreases continuously with increasing pH due to the

6 rapid deprotonation of UDMH, leading to a continuous weakening of the electrostatic attraction.

1
7 The adsorption of UDMH on FCZ@C-600 could be influenced by Van der Waals interactions,

1 1
8 considering the porosity and specific surface area [24]. Moreover, the diffusion process is another

9 factor controlling the adsorption performance. Pyrolysis leads to a larger pore volume and diameter,

1
10 which enhances the diffusion of UDMH molecules within the particles that can reach the internal active

1
11 sites of the adsorbents. As discussed in Section 3.3.1, intraparticle diffusion is the primary rate-limiting

12 step of the diffusion model.

13 Hydrogen and coordination bonds contribute to the adsorption of organic molecules [32,34].

14 FCZ@C-600 contains electronegative N and O atoms from the linker (2-methylimidazole) or the

1 1
15 carbon source (CS). Thus, there are two types of hydrogen bonds: that between FCZ@C-600 as a

16 hydrogen donor (−OH) and UDMH as a hydrogen acceptor (N atoms) and that between FCZ@C-600

1 1
17 as a hydrogen acceptor (N and O atoms) and UDMH as a hydrogen donor (−NH2). Section 3.4.2

18 describes ZIF-8 as a Zn-based MOF with electron-deficient Zn sites interacting with N atoms in

19 UDMH-bearing nucleophilic sites.

20 Thus, the adsorption mechanism of UDMH on FCZ@C-600 combines physisorption and

6
21 chemisorption. Electrostatic interactions (depending on the pH of the solution), diffusion within the

14
1 particles, and hydrogen coordination interactions contribute to the adsorption of UDMH. As shown in

2 Fig. 10, the possible mechanisms include electrostatic interactions, diffusion (first step), and

3 hydrogen/coordination bonds between UDMH and FCZ@C-600 (second step).


4
5 Fig. 910. Potential interaction mechanisms between UDMH and FCZ@C-600 (hydrogen and coordination bonds are shown
6 by dotted yellow and red lines, respectively).

15

You might also like