Toward Sustainable Li-Ion Battery Recycling Green Metal Organic

You might also like

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

pubs.acs.

org/journal/ascecg Research Article

Toward Sustainable Li-Ion Battery Recycling: Green Metal−Organic


Framework as a Molecular Sieve for the Selective Separation of
Cobalt and Nickel
Jędrzej Piątek, Tetyana M. Budnyak, Susanna Monti, Giovanni Barcaro, Robin Gueret,
Erik Svensson Grape, Aleksander Jaworski, A. Ken Inge, Bruno V. M. Rodrigues, and Adam Slabon*
Cite This: ACS Sustainable Chem. Eng. 2021, 9, 9770−9778 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: The growing demand for Li-ion batteries (LIBs)


Downloaded via 36.72.213.194 on November 12, 2022 at 10:52:50 (UTC).

has made their postconsumer recycling an imperative need toward


the recovery of valuable metals, such as cobalt and nickel.
Nevertheless, their recovery and separation from active cathode
materials in LIBs, via an efficient and environmentally friendly
process, have remained a challenge. In this work, we approach a
simple and green method for the selective separation of nickel ions
from mixed cobalt−nickel aqueous solutions under mild
conditions. We discovered that the bioinspired microporous
metal−organic framework (MOF) SU-101 is a selective sorbent
toward Ni2+ ions at pH 5−7 but does not adsorb Co2+ ions.
According to the Freundlich isotherm, the adsorption capacity
toward Ni2+ reached 100.9 mg·g−1, while a near-zero adsorption
capacity was found for Co2+ ions. Ni2+ removal from aqueous solutions was performed under mild conditions (22 °C and pH 5),
with a high yield up to 96%. The presence of Ni2+ ions adsorbed on the surface of the material has been proven by solid-state 1H
nuclear magnetic resonance spectroscopy. Finally, the separation of Ni2+ from Co2+ from binary solutions was obtained with
approximately 30% yield for Ni2+, with a near-zero adsorption of Co2+, which has been demonstrated by UV−vis spectroscopy. The
ion adsorption process of Ni2+ and Co2+ ions was additionally studied by means of classical molecular dynamics calculations (force
fields), which showed that the Ni2+ ions were more prone to enter the MOF canals by replacing some of their coordinated water
molecules. These results offer a green pathway toward the recycling and separation of valuable metals from cobalt-containing LIBs
while providing a sustainable route for waste valorization in a circular economy.
KEYWORDS: adsorption, metal−organic framework, battery recycling, nickel recovery, cobalt recovery

■ INTRODUCTION
Cobalt and nickel are one of the most common components of
for the natural environment.4,5 Furthermore, due to the fact
that the battery market is still expanding and new types of
cathode materials in lithium-ion batteries (LIBs) in the form of batteries have been developed, for example, Al-based batteries,
lithium−metal oxides. These metals are found together in recycling has become even more crucial.6,7
lithium nickel manganese cobalt oxide (NMC) or lithium Sustainable methods for the recycling of LIBs have recently
nickel cobalt aluminum oxide (NCA) cathodes.1,2 The received high interest. However, conventional industrial
European Commission predicts that the demand for cobalt processes still implement non-green methods, that is, high-
supply related to LIBs production will increase 5 times in 2030 temperature processes and toxic chemicals, and do not ensure
and 15 times in 2050, compared to the current supply to EU the same purity grade as the starting materials for the synthesis
countries. Cobalt has already been classified as a critical raw of active electrode materials. In addition, these processes often
material, while nickel is under observation due to the lead to the formation of waste by-products, which in turn may
increasing requisition of LIBs for energy storage and electric
vehicle batteries.3 Depletion of natural deposits of cobalt and Received: March 29, 2021
nickel may result in a global shortage for future prospects. Revised: June 26, 2021
Hence, the recovery of these elements is currently of high Published: July 14, 2021
importance. The majority of electronic wastes containing many
precious elements is, however, not recycled, while the
permeation of toxic elements may have direct consequences
© 2021 The Authors. Published by
American Chemical Society https://doi.org/10.1021/acssuschemeng.1c02146
9770 ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

present hazardous effects to the natural environment.8,9 After cobalt(II) nitrate hexahydrate (98%), and bismuth acetate (99%)
proper mechanical processing, for example, discharging and were obtained from Alfa Aesar, USA. Ethanol absolute, sodium
dismantling, NMC and NCA cathode materials are usually chloride (min 99.8%), sodium acetate (99.8%), hydrochloric acid
treated with mineral acids and separated from the solid fraction (37%), nitric acid (65%), and ammonia (25%) were obtained from
VWR, USA. Sodium hydroxide (98−100.5%) and toluene (≥99.7%)
by filtration. The leachates obtained in these processes were obtained from Honeywell, USA. Iodine solution (0.05 M) and
comprise mixtures of cobalt, nickel, and manganese or bismuth oxide (>98%) were obtained from Merck, USA.
aluminum salts. The separation of these ions is generally Methods. Synthesis. The preparation procedure and character-
performed via hydrometallurgical methods, such as solvent ization of Bi2O(H2O)2(C14H2O8)·nH2 O (SU-101) MOF are
extraction,10−12 precipitation,13,14 and ion exchange.15,16 These described in detail in a previous work.41 Briefly, EA and bismuth
techniques usually implement highly toxic organic solvents acetate were added to a 6% acetic acid mixture with water and stirred
while operating under harsh conditions and requiring addi- at room temperature for 48 h. Next, the suspension was centrifuged at
tional separation steps. In summary, the current existing 8000 rpm for 10 min and dried overnight in a circulating oven (60
methods in industrial processes do not follow the principles of °C). A fine powder of SU-101 was obtained with a yield of 76%. By
using the same synthesis conditions, we have also succeeded with the
green chemistry while also being highly time-consuming.17,18 synthesis of SU-101 via a significantly cheaper substrate, bismuth
Adsorption is a highly efficient surface process to remove oxide (more information about costs can be found in the Supporting
target components from solutions. This method predominates Information). The pH of point zero charge (pHpzc) for SU-101 has
over the aforementioned alternatives due to its capacity to been investigated via the pH drift method using 0.01 M NaCl
operate under mild conditions, therefore minimizing the solutions with pH ranging from 2 to 12. To 5 mL of each solution, 20
implementation of bulk chemicals.19 This method can be mg of SU-101 was suspended, shaken for 24 h, and filtered, after
applied for the removal of metal ions,20 dyes,21 and organic which the final pH was measured. Powder X-ray diffraction (PXRD)
compounds22 from wastewaters, with high effectiveness even measurements were performed using a PANalytical X’Pert PRO
powder diffractometer (Panalytical, Great Britain), with the Cu Kα1,2
for very low concentrations. Adsorption has also been
radiation (λ1 = 1.5406 Å, λ2 = 1.5444 Å). Brightfield images were
implemented as a step process in cobalt ion recovery from obtained using 0.5 s of exposure time on a JEOL JEM-2100 LaB6
LIBs.23 Although many different sorbents have been applied transmission electron microscope (JEOL, Japan) operating at 200 kV,
for cobalt and nickel recovery,24−27 most fail in terms of equipped with a Gatan Orius 200D detector.
selectivity. Thus, these sorbents still need to undergo Adsorption Studies. Batch adsorption experiments were conducted
additional processes in order to separate the metals ions, by suspending 0.1 g of SU-101 in 25 mL of cobalt(II) nitrate or
whenever this is possible. In this context, the development of nickel(II) nitrate solutions in 100 mL flasks, which were shaken for a
sustainable methods for the selective separation of metal ions given time in a Heidolph Unimax 1010 incubating shaker (Germany)
from water solutions has become crucial and urgent. at 180 rpm and 22 °C. The pH influence on adsorption was evaluated
by preparing solutions containing 50 mg·L−1 Co2+ or Ni2+ and
Metal−organic frameworks (MOFs) are hybrid porous
adjusting the pH from 2 to 8 with 0.01 M HNO3 and 0.01 M NH4OH
materials comprising metal ions or clusters and organic ligands. solutions. Isotherms have been evaluated with solutions of initial
Over the past two decades, these materials gained much concentration of 2−80 mg·L−1 at pH 2.0 and 5.0 for Co2+ and a
scientific interest due to their broad range of application, concentration of 2−520 mg·L−1 at pH 5.0 and 7.0 for Ni2+ ions. The
including catalysis, hydrogen storage, carbon capture, semi- equilibrium time needed for the adsorption was investigated by
conductors, drug delivery systems, and biological imaging and suspending the sorbent in 50 and 100 mg·L−1 Co2+ or Ni2+ solutions
sensing.28−35 As sorbents, they have been investigated for the for 0.25−24 h. The final concentration of metal ions was determined
adsorption of dyes,36 H2,37 O2,38 and CO239 gases, as well as by the methods described by Marchenko42 using a UV-3100PC
heavy metals40 from aqueous solutions. Recently, we have spectrophotometer (VWR, USA). Absorbance was measured by the
formation of Co2+ complexes with 4-(2′-pyridylazo)resorcinol (at 500
reported on the green synthesis of a bismuth ellagate MOF by nm) and Ni2+ complexes with dimethylglyoxime (at 470 nm). The
using nonhazardous chemicals in a simple synthesis process adsorption capacity (qeq) was calculated using eq 1
under ambient conditions. Besides a high chemical stability,
this MOF presents interesting physico-chemical and surface (C0 − Ceq)V
properties, which places it as a material with a great potential qeq =
m (1)
toward heavy metal ion recycling.41
In this work, bismuth ellagate Bi2O(H2O)2(C14H2O8)·nH2O where C0 is the initial metal concentration (mg·L−1), Ceq is the
equilibrium metal concentration (mg·L−1), V is the sample volume
MOF (SU-101) has been investigated as a selective agent for
(L), and m is the sorbent mass (g). To investigate the mechanism
Ni2+ removal from mixed cobalt−nickel aqueous solutions. behind the adsorption of Ni2+ ions, three isotherm models were
Ni2+ adsorption kinetics have been determined using pseudo- determined: Langmuir, Freundlich, and Temkin. In the Langmuir
first- and pseudo-second-order reaction equations. The model, the adsorption occurs on a homogeneous surface, and
possibility of intraparticle diffusion has been verified, while adsorbed molecules do not interact with each other. The Langmuir
Langmuir, Freundlich, and Temkin models have been applied model can be expressed as follows
to investigate the isotherm model of Ni2+ adsorption. This
Ceq Ceq 1
green MOF achieves Co−Ni separation at room temperature, = +
yielding the opportunity to recover cobalt as the most qeq q0 KLq0 (2)
important metal from cathode materials in LIBs. −1


where Ceq is the equilibrium concentration of metal ions (mg·L ), qeq
is the amount of the adsorbed ions (mg·g−1), q0 is the sorption
EXPERIMENTAL SECTION capacity (mg·g−1), and KL is the equilibrium constant (L·mg−1). The
Materials. 1-Nitroso-2-naphthol-3,6-disulfonic acid sodium salt Langmuir isotherm can also be expressed with a dimensionless
hydrate (pure, indicator grade), sodium tetraborate (98%), ellagic separating factor RL
acid (EA) (97%), acetic acid (99.7%), 4-(2′-pyridylazo)resorcinol 1
(97+%, ACS), and dimethylglyoxime (99+%) were obtained from RL =
Acros Organics, Belgium. Nickel(II) nitrate hexahydrate (98%), 1 + KLC0 (3)

9771 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

where C0 is the initial concentration of the adsorbate (mg·g−1).


The Freundlich model describes multilayer adsorptions and can be
expressed as
1
log qeq = log KF + log Ceq
n (4)
where KF and n are the Freundlich constants of sorption capacity (L·
mg−1) and sorption intensity, respectively.
The Temkin model has been calculated with the following equation
RT RT
CS = ln KT + ln Ceq
bT bT (5)
where CS is the concentration of metal in the solid phase (mol·g−1),
KT is the model constant (L·g−1), R is the gas constant (8.314 J·
mol−1·K−1), T is the absolute temperature (K), bT is the heat of
adsorption (J·mol−1), and Ceq is the equilibrium metal concentration Figure 1. SU-101 structure with hydrogen atoms and water molecules
in the aqueous phase (mol·L−1).25,43 in the pores. Bi3+, C, O, and H atoms are purple, cyan, red, and white,
The kinetics of Ni2+ adsorption have been investigated using respectively. Unit cell (green lines).
pseudo-first- and pseudo-second-order models. The intraparticle
diffusion of the adsorption has also been verified. Pseudo-first- and tion with the Perdew−Burke−Ernzerhof functional.50 Periodic
pseudo-second-order models were calculated using the following boundary conditions were applied in all directions, and a cutoff
equations. energy of 600 eV was employed to obtain convergent results. This
Pseudo-first-order equation model was used to create a larger system to simulate the ion
k1t adsorption process through classical molecular dynamics (MD)
log(qeq − qt ) log qeq − calculations (force fields).
2.303 (6)
The dynamics of adsorption for nickel and cobalt ions in the MOF
Pseudo-second-order equation at an ambient temperature have been explored by means of reactive
MD simulations (ReaxFF). The force field parameters were extracted
t 1 t from the data contained in the ADF/ReaxFF package,51 integrating
= +
qt k 2qeq 2 qeq (7) the CHONSMgPNaCuCl_v2.ff force field parameters,52 which were
appropriate for the organic acid, with the Bi, Ni, and Co parameters
−1
where qeq and qt are the adsorption capacities (mg·g ) at equilibrium extracted from force fields belonging to the water branch. A more
and at any instant time of t, respectively, k1 is the rate constant of the extended model (supercell) was prepared by replicating the original
pseudo-first-order reaction (1·min−1), and k2 is the rate constant of crystal cell, containing the pore waters, three times in x and y and five
the pseudo-second-order reaction (g·mg−1·min−1).25,44 times in z directions (55.865 × 55.865 × 27.733 Å3). The metal ion
The Weber and Morris equation has been applied to examine adsorption process was simulated by extending the z box size to
whether intraparticle diffusion occurs 40.000 Å and adding a water layer (approximately 12 Å thick) with
around 1000 water molecules, 50 metal ions (Ni2+ and Co2+two
qt = KIPDt 0.5 + C (8) different simulations), and 100 OH− as counterions, randomly
−1 −0.5
where KIPD is the intraparticle diffusion rate (mg·g ·min ) and C is distributed in the solution (Figure 2).
a constant.25,45
The separation of Ni2+ from Co2+ via adsorption was evaluated by
stirring mixed cobalt−nickel solutions, which contained 4 and 10 mg·
L−1 of each metal, for 24 h with 0.1 g of the SU-101 sorbent at 22 °C.
After that, solutions were filtered, and the final concentration of Ni2+
and Co2+ was determined. For this, we added 2 mL of samples in 25
mL flasks, to which 5 mL of 0.5 mmol·L−1 nitroso-2-naphthol-3,6-
disulfonic acid sodium salt and 7.5 mL of acetate buffer (pH 5.5) were
added and filled with water up to 25 mL. Then, UV−vis spectra of the
samples were recorded from 800 to 350 nm using a UV−vis
spectrophotometer. Simultaneous determination of these two metal
ions in this study was based on the method proposed by Zhou et al.46
The 1H magic angle spinning (MAS) nuclear magnetic resonance
(NMR) experiments were performed at the magnetic field B0 = 14.1 T
(Larmor frequency of 600.12 MHz) and MAS rate vr = 60.00 kHz on
a Bruker AVANCE-III spectrometer equipped with a 1.3 mm MAS
probehead. The 1H acquisitions involved a rotor-synchronized,
double-adiabatic spin-echo sequence with 90° 1.25 μs excitation
pulse, followed by two 50.0 μs tanh/tan high-power adiabatic pulses
(SHAPs) with a 5 MHz frequency sweep.47,48 All pulses operated at
the nutation frequency vnut = 200 kHz. 128 signal transients with a 5 s
relaxation delay were accumulated for each spectrum. Shifts were
referenced with respect to neat tetramethylsilane.
Computational Methods. The atomic coordinates of SU-101 and
unit cell parameters (18.6217 × 18.6217 × 5.5466 Å3Figure 1) Figure 2. Starting structure of the MOF model used in the
were obtained from the Supporting Information of Grape et al.41 An simulations depicted through its solvent-accessible surface. Pore
optimization was performed at the density functional theory level after waters are red spheres (oxygen), whereas the added water layer and
adding the hydrogen atoms using the Quantum Espresso package,49 Ni2+ ions are shown as red-white wires (O, H) and green spheres.
ultrasoft pseudopotentials, and the generalized gradient approxima- Top: x−y plane. Bottom: sections in the z-direction.

9772 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

The whole system was energy-minimized, gradually heated to 300 Ni2+. Additional experiments at pH 2.0 for Co2+ and 7.0 for
K, and equilibrated at that temperature for about 50 ps. Then, Ni2+ were also carried out for the isolated ions.
production MD simulations were carried out in the NVT ensemble for Isotherms for Co2+ adsorption were measured at pH 2.0 and
a time sufficient to observe the adsorption (more than 1 ns). The 5.0, at 22 °C, for 3 h (Figure S4 in the Supporting
system structures were collected every 0.1 ps. The temperature was
controlled through the Nosé−Hoover thermostat with a relaxation
Information). Poor sorption capacities at both pH values
constant of 0.05 ps, and the time step was set to 0.2 fs. The analysis of indicate that SU-101 is not suitable to remove Co2+ ions from
the production trajectories was focused on the positions of the ions aqueous solutions. Conversely, a high sorption capacity of
inside the MOF. 100.9 mg·g−1 (87.3% efficiency) was found for Ni2+ at pH 5.0;

■ RESULTS AND DISCUSSION


Adsorption Study on Ni2+ and Co2+ Ions. In this work,
at pH 7.0, the maximal capacity decreased to 23.3 mg·g−1 after
3 h (Figure 4). In the literature, one can find numerous reports

we have performed all experiments using SU-101 synthesized


using bismuth acetate as the precursor. The main character-
izations of this material have been reported in a previous work,
and it has been found that SU-101 is chemically stable in a pH
range from 2 to 14.41 We have also performed the synthesis of
SU-101 using a much cheaper bismuth source (bismuth
oxide). PXRD measurements revealed the same final structure
(Figure S1 in the Supporting Information), while transmission
electron microscopy (TEM) micrographs revealed the
formation of polydispersed crystals (Figure S2 in the
Supporting Information). The pHpzc value reveals at which
pH the surface of the material is neutral, that is, same number
of positive and negative charges. When the pH is below pHpzc,
the material’s surface is positively charged, whereas if the pH is
above pHpzc, the surface is negatively charged.53 The pH of
point zero charge for SU-101 was found to be 2.29 (Figure S3
in the Supporting Information), which means that pH values
above pHpzc favor the adsorption of positively charged metal
ions. Initially, batch adsorption experiments with respect to pH
were conducted (Figure 3). For Co2+ ions, the sorption
Figure 4. Adsorption isotherms of nickel ions on SU-101.

on sorbents for Ni2+ ions with higher maximal capacity, for


example, modified silica (172.4 mg·g−1),25 activated carbon
(140.85 mg·g−1),54 or algae (181.2 mg·g−1).55 MOFs have
been long recognized by their high efficiency for heavy metal
ion removal,56,57 although their selectivity toward specific
elements is not often the object of study. Conversely, SU-101
has demonstrated a notable selectivity toward Ni2+ ions over
Co2+ ions. Isotherm models have been calculated for Ni2+
adsorption at pH 5.0, and the parameters are shown in Table 1
and Figures S5−S7 (Supporting Information). The Freundlich
isotherm model was found to be the most suitable based on
the experimental data, with the highest correlation coefficient
(R2 = 0.988). The Freundlich model assumes that the
adsorption mechanism occurs on a heterogeneous surface of

Table 1. Langmuir, Freundlich, and Temkin Isotherm


Figure 3. Effect of initial pH on the adsorption of cobalt and nickel Models’ Parameters of Ni2+ Adsorption at pH 5.0
ions.
isotherm model parameter result ± SD
Langmuir q0 (mg·g−1) 116.29 ± 0.33
capacity was near-zero in almost the entire pH range, while for KL (L·mg−1) 0.0058 ± 0.001
nickel(II) ions, the highest capacity was found at pH 8.0 (14.6 RL 0.526 ± 0.26
mg·g−1, 80.5%). However the most interesting results were R2 0.548
obtained at pH 5.0, where Ni2+ sorption starts to equilibrate, Freundlich KF (L·mg−1) 1.833 ± 0.37
reaching 14.2 mg·g −1 (78.9% efficiency), while Co 2+ n 0.852 ± 0.21
adsorption is near-zero (0.08 mg·g−1, 0.4% efficiency). Due R2 0.988
to the selectivity for Ni2+ sorption over Co2+ ions, further Temkin bT (J·mol−1) 80.192 × 103 ± 1.33
studies at pH 5.0 were carried out in mixed cobalt−nickel KT (J·mol−1) 3.049 × 1025 ± 2.87
solutions in order to verify the selectivity of SU-101 toward R2 0.713

9773 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

a sorbent, with possible interactions between adsorbed ions or


molecules.43
The kinetics for Ni2+ adsorption was investigated with initial
concentrations of 50 and 100 mg·L−1 at pH 5.0 (Figure 5).

Figure 6. Pseudo-second-order plot of Ni2+ adsorption.

suitable for the adsorption of Ni2+ on SU-101. This result also


shows that chemisorption is the main mechanism behind Ni2+
adsorption.44 The plot qt versus time (t0.5) is nonlinear in both
cases, and it indicates that strong interactions and boundary
layer diffusion may control the rate of adsorption (Figure S9 in
Figure 5. Adsorption kinetics of nickel ions on SU-101.
the Supporting Information).
In Figure 7a, 1H MAS NMR spectra collected from SU-101
The adsorption of nickel ions by SU-101 is a rapid and efficient samples before and after adsorption are presented. In the
process, even after 15 min. The highest efficiency was reached spectrum of the as-synthesized sample (black trace), 1H signals
after 24 h of shaking (93.3 and 96.0% for 50 and 100 mg·L−1, from EA moieties appear at 7 and 3 ppm and correspond to
respectively), although it can be observed that the curves reach C−H and O−H protons, respectively. A broad signal centered
an equilibrium after 4−6 h of the process (87−92% removal at 5 ppm originates from physisorbed water. Yet, another
efficiency in both cases). The parameters for all kinetic models broad resonance at 10 ppm results from COOH groups, which
can be found in Table 2. A pseudo-first-order model was along with signals from CH3 groups at ∼1 ppm indicates the
presence of acetic acid remaining in the material after
synthesis. A sharp signal at 0.5 ppm can be attributed to
Table 2. Kinetic Parameters of Ni2+ Adsorption on SU-101
hydroxyl groups associated with inorganic building units. The
results spectrum collected from the sample after adsorption (red
kinetic model parameter 50 ppm 100 ppm trace) reveals an increased signal intensity from H2O at 4.7
pseudo-first order qeq,exp (mg·g−1) 12.11 23.85
ppm and substantially reduced resonances of acetic acid at 10
qeq,cal (mg·g−1) 2.71 4.88
and ∼1 ppm. Noteworthy, O−H protons from EA moieties, as
k1 (1·min−1) 0.048 0.072
well as those involved in hydroxyl groups, are almost
R2 0.515 0.615
completely gone. Based on these observations, it can be
pseudo-second qeq,exp (mg·g−1) 12.11 23.85
assumed that the process of Ni2+-ion adsorption on SU-101
order can be regarded in terms of an ion exchange with labile protons
qeq,cal (mg·g−1) 12.28 24.09 present in the MOF structure (Figure 7b).
k2 (g·mg−1·min−1) 2.73 × 10−3 1.74 × 10−3 Adsorption Mechanism on SU-101. From the theoreti-
R2 0.998 0.999 cal inspection of the trajectories, it was apparent that the Co2+
intraparticle KIPD 0.104 0.212 ions had the tendency to remain at the MOF−water interface
diffusion (mg·g−1·min−0.5) and eventually adsorb into the MOF model just below the
C 8.366 16.78 MOF−water interface (around 3 Å) through their interposing
R2 0.667 0.581 water layer, which was well preserved. Conversely, the Ni2+
ions were more prone to enter the MOF channels by replacing
constructed by fitting the experimental data from the slope and some of their coordinated water molecules with those in the
intercept of the log (qeq − qt) versus t plot (Figure S8 in the pores and engaging in interactions with the carbonyl oxygen
Supporting Information). This model was found to be not from the EA molecules. This is confirmed by examining the
suitable for Ni2+ adsorption because the calculated equilibrium number of adsorbed ions as a function of the simulation time,
capacities in both cases did not match the experimental values as shown in Figure 8.
while also having a low correlation coefficient. Instead, a The adsorption is also confirmed by the presence of sharp
pseudo-second-order model was constructed by plotting t/qt peaks at short distances in the radial distribution functions
versus t (Figure 6). It was found that the calculated values of between the oxygen atom MOF and ions (Figure 9).
qeq were similar to the ones obtained experimentally and the Examining the RDF plots, it can be noticed that Co2+ could
correlation coefficient in both cases was also high (∼1.0), be close to O(EA) at the surface but not inside the MOF near
revealing the pseudo-second-order reaction model as the most the oxygen connected to Bi and well solvated by water (bulk).
9774 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 7. (a) 1H MAS NMR spectra of SU-101 before (black) and after (red) Ni ion adsorption. The data were collected at 14.1 T and 60.00 kHz
MAS rate. (b) Proposed mechanism for Ni2+ adsorption on SU-101 from a mixed nickel−cobalt solution.

Figure 9. Normalized RDFs of the metal ions with three types of


oxygen atoms, namely, carbonyl oxygen of EA [O(EA)], oxygen
atoms connected to Bi [O(Bi)], and oxygen atoms of the water
Figure 8. Top: number of ions adsorbed inside the MOF channels. molecules not adsorbed in the MOF [O(water)]. Representative
Bottom: percentage probability of contacts with the O(CO) oxygen connection/entrapment of the Co ion surrounded by water at the
of EA. MOF−water interface.

Instead, nickel ions are found in solution and in close contact traces, respectively), as well as reference solutions containing 4
with both O(Bi) and O(EA) inside the framework (Figure 10). and 10 mg·L−1 Co2+ (blue trace) and 4 or 10 mg·L−1 Ni2+
The simultaneous analysis of nickel and cobalt ions is (green trace). In case of 100% recovery of Ni2+, we should
challenging, even by spectrophotometric methods. Herein, we expect the spectrum after adsorption to overlap with the
used a method recently reported by Zhou et al.46 for the spectrum of a pure Co2+ solution.
determination of Ni2+ and Co2+ in solutions using the UV−vis By analyzing the spectra, we can observe that the solution
technique. Our main goal was to evaluate the selectivity of the after adsorption has a significant decrease in the band intensity
MOF, SU-101, toward Ni2+ ions over Co2+ in mixed cobalt− in the region from 500 to 450 nm, which corresponds to nickel
nickel aqueous solutions. Figures 11 and S10 present the ions (green trace). The presence of nickel ions broadens the
spectra of solutions before and after adsorption (black and red Co2+ spectrum, and narrowing of the band from 500 to 450 nm
9775 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

adsorption of these ions, and hence, we conclude that there


was near-zero adsorption of Co2+. These results confirmed the
selectivity of SU-101 toward the separation of Ni2+ from mixed
cobalt−nickel solutions.

■ CONCLUSIONS
In this work, we demonstrate the intriguing sorption selectivity
of a bioinspired microporous bismuth ellagate MOF, SU-101,
for the separation of Ni2+ from mixed cobalt−nickel aqueous
solutions. A near-zero sorption of Co2+ was observed in almost
the entire pH range examined, while the adsorption capacity of
Ni2+ was found to be competitively high (up to 100.9 mg·g−1).
The evaluation of isotherms showed that the adsorption of
nickel ions follows the Freundlich model, pointing out that the
sorption occurs on a heterogeneous surface. Fast and efficient
Ni2+ adsorption (up to 96%) was obtained following the
pseudo-second-order kinetic model, meaning that chemisorp-
tion is the main mechanism governing the reaction. Finally, the
successful separation of Ni2+ from mixed cobalt−nickel
aqueous solutions was performed, reaching approximately
30% of Ni2+ recovery and near-zero Co2+ recovery after 24 h.
SU-101 revealed the ability of separating those two ions from
aqueous solutions, which provides a great prospect for future
applications in spent LIBs. The ion adsorption process of Ni2+
and Co2+ ion was additionally studied by means of classical
Figure 10. (Top) Snapshot from the Ni2+ MD where two ions have MD calculations (force fields), which showed that the Ni2+
reached the middle of the channels, whereas other adsorbed ions are ions were more prone to enter the MOF canals by replacing
relatively close to the openings. (Bottom) Binding mode of Ni2+ to some of their coordinated water molecules. These results offer
the carbonyl oxygen of EA. Bi (purple spheres). a straightforward pathway toward the recycling and separation
of valuable metals while providing a sustainable route for waste
valorization in a circular economy.

■ ASSOCIATED CONTENT
* Supporting Information

The Supporting Information is available free of charge at


https://pubs.acs.org/doi/10.1021/acssuschemeng.1c02146.
Cost comparison of SU-101 synthesis using different
substrates, PXRD patterns of SU-101, TEM micrograph
of SU-101, graph of the pH of point zero charge, Co2+
adsorption isotherms, determination of isotherm models
for Ni2+ adsorption, determination of kinetic parameters
for Ni2+ adsorption, and UV−vis spectra of Co2+/Ni2+
mixed solutions before and after adsorption (PDF)

■ AUTHOR INFORMATION
Corresponding Author
Figure 11. UV−vis spectra of the following: initial solution containing
Adam Slabon − Department of Materials and Environmental
10 mg·L−1 Ni2+ and Co2+ ions before (black trace) and after (red
trace) adsorption, 10 mg·L−1 Co2+ solution (blue trace), and 10 mg· Chemistry, Stockholm University, 106 91 Stockholm,
L−1 Ni2+ solution (green trace). Sweden; orcid.org/0000-0002-4452-1831;
Email: adam.slabon@mmk.su.se

indicates the decrease of the initial concentration of nickel ions Authors


in the solution. Based on the obtained data, we calculated a Jędrzej Piątek − Department of Materials and Environmental
Ni2+ recovery of approximately 30% with respect to the initial Chemistry, Stockholm University, 106 91 Stockholm, Sweden
solution. Due to the overlap of nickel and cobalt absorption Tetyana M. Budnyak − Department of Materials and
bands in the range from 450 to 400 nm, the slight decrease in Environmental Chemistry, Stockholm University, 106 91
the intensity of the maximum absorption corresponding to Stockholm, Sweden; orcid.org/0000-0003-2112-9308
cobalt ions after adsorption is also related to the decrease of Susanna Monti − CNR-ICCOMInstitute of Chemistry of
Ni2+ concentration. There are no significant changes in the Organometallic Compounds, 56124 Pisa, Italy; orcid.org/
Co2+ spectrum after adsorption that would point out the 0000-0002-3419-7118
9776 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Giovanni Barcaro − CNR-IPCFInstitute for Chemical and Performance Al Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12,
Physical Processes, 56124 Pisa, Italy; orcid.org/0000- 2572−2580.
0002-5520-5914 (8) Yin, H.; Xing, P. Pyrometallurgical Routes for the Recycling of Spent
Robin Gueret − Department of Materials and Environmental Lithium-Ion Batteries BTRecycling of Spent Lithium-Ion Batteries:
Chemistry, Stockholm University, 106 91 Stockholm, Sweden Processing Methods and Environmental Impacts; An, L., Ed.; Springer
International Publishing: Cham, 2019; pp 57−83.
Erik Svensson Grape − Department of Materials and
(9) Zheng, X.; Zhu, Z.; Lin, X.; Zhang, Y.; He, Y.; Cao, H.; Sun, Z. A
Environmental Chemistry, Stockholm University, 106 91 Mini-Review on Metal Recycling from Spent Lithium Ion Batteries.
Stockholm, Sweden; orcid.org/0000-0002-8956-5897 Engineering 2018, 4, 361−370.
Aleksander Jaworski − Department of Materials and (10) Chen, W.-S.; Ho, H.-J. Recovery of Valuable Metals from
Environmental Chemistry, Stockholm University, 106 91 Lithium-Ion Batteries NMC Cathode Waste Materials by Hydro-
Stockholm, Sweden; orcid.org/0000-0002-7156-559X metallurgical Methods. Metals 2018, 8, 321.
A. Ken Inge − Department of Materials and Environmental (11) Tait, B. K. Cobalt-Nickel Separation: The Extraction of
Chemistry, Stockholm University, 106 91 Stockholm, Cobalt(II) and Nickel(II) by Cyanex 301, Cyanex 302 and Cyanex
Sweden; orcid.org/0000-0001-9118-1342 272. Hydrometallurgy 1993, 32, 365−372.
Bruno V. M. Rodrigues − Department of Materials and (12) Wellens, S.; Thijs, B.; Möller, C.; Binnemans, K. Separation of
Environmental Chemistry, Stockholm University, 106 91 cobalt and nickel by solvent extraction with two mutually immiscible
Stockholm, Sweden ionic liquids. Phys. Chem. Chem. Phys. 2013, 15, 9663−9669.
(13) Joulié, M.; Laucournet, R.; Billy, E. Hydrometallurgical Process
Complete contact information is available at: for the Recovery of High Value Metals from Spent Lithium Nickel
https://pubs.acs.org/10.1021/acssuschemeng.1c02146 Cobalt Aluminum Oxide Based Lithium-Ion Batteries. J. Power Sources
2014, 247, 551−555.
Author Contributions (14) Chen, X.; Zhou, T.; Kong, J.; Fang, H.; Chen, Y. Separation and
All authors have given approval to the final version of the Recovery of Metal Values from Leach Liquor of Waste Lithium Nickel
manuscript. Cobalt Manganese Oxide Based Cathodes. Sep. Purif. Technol. 2015,
141, 76−83.
Notes (15) Grinstead, R. R. Selective Absorption of Copper, Nickel, Cobalt
The authors declare no competing financial interest. and Other Transition Metal Ions from Sulfuric Acid Solutions with


the Chelating Ion Exchange Resin Xfs 4195. Hydrometallurgy 1984,
ACKNOWLEDGMENTS 12, 387−400.
(16) Botelho Junior, A. B.; Dreisinger, D. B.; Espinosa, D. C. R. A
This work was financially supported by the Stiftelsen Olle Review of Nickel, Copper, and Cobalt Recovery by Chelating Ion
Engkvist (project no. 198-0329). E.S.G. and A.K.I. acknowl- Exchange Resins from Mining Processes and Mining Tailings. Mining,
edge support from the Swedish Foundation for Strategic Metall. Explor. 2019, 36, 199−213.
Research (SSF). J.P. would like to thank Jianhong Chen for (17) Anastas, P.; Eghbali, N. Green Chemistry: Principles and
help and discussion. Practice. Chem. Soc. Rev. 2010, 39, 301−312.


(18) Anastas, P. T.; Warner, J. C. Green Chemistry: Theory and
REFERENCES Practice; Oxford University Press: New York, 1998.
(19) Dąbrowski, A. Adsorptionfrom Theory to Practice. Adv.
(1) Winslow, K. M.; Laux, S. J.; Townsend, T. G. A Review on the Colloid Interface Sci. 2001, 93, 135−224.
Growing Concern and Potential Management Strategies of Waste (20) Ubando, A. T.; Africa, A. D. M.; Maniquiz-Redillas, M. C.;
Lithium-Ion Batteries. Resour., Conserv. Recycl. 2018, 129, 263−277. Culaba, A. B.; Chen, W.-H.; Chang, J.-S. Microalgal Biosorption of
(2) Wang, H.; Huang, K.; Zhang, Y.; Chen, X.; Jin, W.; Zheng, S.; Heavy Metals: A Comprehensive Bibliometric Review. J. Hazard.
Zhang, Y.; Li, P. Recovery of Lithium, Nickel, and Cobalt from Spent Mater. 2021, 402, 123431.
Lithium-Ion Battery Powders by Selective Ammonia Leaching and an (21) Budnyak, T. M.; Aminzadeh, S.; Pylypchuk, I. V.; Sternik, D.;
Adsorption Separation System. ACS Sustainable Chem. Eng. 2017, 5,
Tertykh, V. A.; Lindström, M. E.; Sevastyanova, O. Methylene Blue
11489−11495.
Dye Sorption by Hybrid Materials from Technical Lignins. J. Environ.
(3) Keersemaker, M. Critical Raw Materials Resilience: Charting a
Chem. Eng. 2018, 6, 4997−5007.
Path towards Greater Security and Sustainability. Communication from
(22) Bergfreund, J.; Bertsch, P.; Fischer, P. Adsorption of Proteins to
the Commission to the European Parliament, the Council, the European
Economic and Social Committee and the Committee of the Regions; Fluid Interfaces: Role of the Hydrophobic Subphase. J. Colloid
European Commission: Brussels, 2020; pp 69−82. Interface Sci. 2021, 584, 411−417.
(4) Grant, K.; Goldizen, F. C.; Sly, P. D.; Brune, M.-N.; Neira, M.; (23) Piątek, J.; Afyon, S.; Budnyak, T. M.; Budnyk, S.; Sipponen, M.
van den Berg, M.; Norman, R. E. Health Consequences of Exposure H.; Slabon, A. Sustainable Li-Ion Batteries: Chemistry and Recycling.
to E-Waste: A Systematic Review. Lancet 2013, 1, e350−e361. Adv. Energy Mater. 2020, 2003456.
(5) de Bruin-Dickason, C.; Budnyk, S.; Piątek, J.; Jenei, I.-Z.; (24) Prado, A. G. S.; Arakaki, L. N. H.; Airoldi, C. Adsorption and
Budnyak, T. M.; Slabon, A. Valorisation of Used Lithium-Ion Separation of Cations on Silica Gel Chemically Modified by
Batteries into Nanostructured Catalysts for Green Hydrogen from Homogeneous and Heterogeneous Routes with the Ethylenimine
Boranes. Mater. Adv. 2020, 1, 2279−2285. Anchored on Thiol Modified Silica Gel. Green Chem. 2002, 4, 42−46.
(6) Wang, D.-Y.; Huang, S.-K.; Liao, H.-J.; Chen, Y.-M.; Wang, S.- (25) Piątek, J.; de Bruin-Dickason, C. N.; Jaworski, A.; Chen, J.;
W.; Kao, Y.-T.; An, J.-Y.; Lee, Y.-C.; Chuang, C.-H.; Huang, Y.-C.; Lu, Budnyak, T.; Slabon, A. Glycine-Functionalized Silica as Sorbent for
Y.-R.; Lin, H.-J.; Chou, H.-L.; Chen, C.-W.; Lai, Y.-H.; Dong, C.-L. Cobalt(II) and Nickel(II) Recovery. Appl. Surf. Sci. 2020, 530,
Insights into Dynamic Molecular Intercalation Mechanism for AlC 147299.
Battery by Operando Synchrotron X-Ray Techniques. Carbon 2019, (26) Ferri, M.; Campisi, S.; Gervasini, A. Nickel and Cobalt
146, 528−534. Adsorption on Hydroxyapatite: A Study for the de-Metalation of
(7) Lee, T.-S.; Patil, S. B.; Kao, Y.-T.; An, J.-Y.; Lee, Y.-C.; Lai, Y.-H.; Electronic Industrial Wastewaters. Adsorption 2019, 25, 649−660.
Chang, C.-K.; Cheng, Y.-S.; Chuang, Y.-C.; Sheu, H.-S.; Wu, C.-H.; (27) Budnyak, T. M.; Modersitzki, S.; Pylypchuk, I. V.; Piątek, J.;
Yang, C.-C.; Cheng, R.-H.; Lee, C.-Y.; Peng, P.-Y.; Lai, L.-H.; Lee, H.- Jaworski, A.; Sevastyanova, O.; Lindström, M. E.; Slabon, A. Tailored
H.; Wang, D.-Y. Real-Time Observation of Anion Reaction in High Hydrophobic/Hydrophilic Lignin Coatings on Mesoporous Silica for

9777 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Sustainable Cobalt(II) Recycling. ACS Sustainable Chem. Eng. 2020, (46) Zhou, F.; Li, C.; Zhu, H.; Li, Y. A Novel Method for
8, 16262−16273. Simultaneous Determination of Zinc, Nickel, Cobalt and Copper
(28) Wu, S.; Liu, J.; Wang, H.; Yan, H. A Review of Performance Based on UV−Vis Spectrometry. Optik 2019, 182, 58−64.
Optimization of MOF-Derived Metal Oxide as Electrode Materials for (47) Kervern, G.; Pintacuda, G.; Emsley, L. Fast Adiabatic Pulses for
Supercapacitors. Int. J. Energy Res. 2019, 43, 697−716. Solid-State NMR of Paramagnetic Systems. Chem. Phys. Lett. 2007,
(29) Gangu, K. K.; Maddila, S.; Mukkamala, S. B.; Jonnalagadda, S. 435, 157−162.
B. Characteristics of MOF, MWCNT and Graphene Containing (48) Hwang, T.-L.; Van Zijl, P. C. M.; Garwood, M. Fast Broadband
Materials for Hydrogen Storage: A Review. J. Energy Chem. 2019, 30, Inversion by Adiabatic Pulses. J. Magn. Reson. 1998, 133, 200−203.
132−144. (49) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.;
(30) Goetjen, T. A.; Liu, J.; Wu, Y.; Sui, J.; Zhang, X.; Hupp, J. T.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni, M.; Dabo,
Farha, O. K. Metal-Organic Framework (MOF) Materials as I.; Dal Corso, A.; de Gironcoli, S.; Fabris, S.; Fratesi, G.; Gebauer, R.;
Polymerization Catalysts: A Review and Recent Advances. Chem. Gerstmann, U.; Gougoussis, C.; Kokalj, A.; Lazzeri, M.; Martin-
Samos, L.; Marzari, N.; Mauri, F.; Mazzarello, R.; Paolini, S.;
Commun. 2020, 56, 10409−10418.
Pasquarello, A.; Paulatto, L.; Sbraccia, C.; Scandolo, S.; Sclauzero, G.;
(31) Ren, J.; Langmi, H. W.; North, B. C.; Mathe, M. Review on
Seitsonen, A. P.; Smogunov, A.; Umari, P.; Wentzcovitch, R. M.
Processing of Metal−Organic Framework (MOF) Materials towards
QUANTUM ESPRESSO: A Modular and Open-Source Software
System Integration for Hydrogen Storage. Int. J. Energy Res. 2015, 39, Project for Quantum Simulations of Materials. J. Phys.: Condens.
607−620. Matter 2009, 21, 395502.
(32) Xuan, W.; Zhu, C.; Liu, Y.; Cui, Y. Mesoporous Metal−Organic (50) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient
Framework Materials. Chem. Soc. Rev. 2012, 41, 1677−1695. Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
(33) Zhu, Q.-L.; Xu, Q. Metal-Organic Framework Composites. (51) te Velde, G.; Bickelhaupt, F. M.; Baerends, E. J.; Fonseca
Chem. Soc. Rev. 2014, 43, 5468−5512. Guerra, C.; van Gisbergen, S. J. A.; Snijders, J. G.; Ziegler, T.
(34) Bagi, S.; Wright, A. M.; Oppenheim, J.; Dincă, M.; Román- Chemistry with ADF. J. Comput. Chem. 2001, 22, 931−967.
Leshkov, Y. Accelerated Synthesis of a Ni2Cl2(BTDD) Metal- (52) Monti, S.; Corozzi, A.; Fristrup, P.; Joshi, K. L.; Shin, Y. K.;
Organic Framework in a Continuous Flow Reactor for Atmospheric Oelschlaeger, P.; van Duin, A. C. T.; Barone, V. Exploring the
Water Capture. ACS Sustainable Chem. Eng. 2021, 9, 3996−4003. Conformational and Reactive Dynamics of Biomolecules in Solution
(35) Xu, W.; Yaghi, O. M. Metal-Organic Frameworks for Water Using an Extended Version of the Glycine Reactive Force Field. Phys.
Harvesting from Air, Anywhere, Anytime. ACS Cent. Sci. 2020, 6, Chem. Chem. Phys. 2013, 15, 15062−15077.
1348−1354. (53) Kołodyńska, D.; Gęca, M.; Pylypchuk, I. V.; Hubicki, Z.
(36) Huang, J.; Huang, D.; Zeng, F.; Ma, L.; Wang, Z. Photocatalytic Development of New Effective Sorbents Based on Nanomagnetite.
MOF Fibrous Membranes for Cyclic Adsorption and Degradation of Nanoscale Res. Lett. 2016, 11, 152−161.
Dyes. J. Mater. Sci. 2021, 56, 3127−3139. (54) Anoop Krishnan, K.; Sreejalekshmi, K. G.; Baiju, R. S.
(37) Ibarra, I. A.; Lin, X.; Yang, S.; Blake, A. J.; Walker, G. S.; Nickel(II) Adsorption onto Biomass Based Activated Carbon
Barnett, S. A.; Allan, D. R.; Champness, N. R.; Hubberstey, P.; Obtained from Sugarcane Bagasse Pith. Bioresour. Technol. 2011,
Schröder, M. Structures and H2 Adsorption Properties of Porous 102, 10239−10247.
Scandium Metal-Organic Frameworks. Chem.Eur. J. 2010, 16, (55) Kalyani, S.; Srinivasa Rao, P.; Krishnaiah, A. Removal of Nickel
13671−13679. (II) from Aqueous Solutions Using Marine Macroalgae as the Sorbing
(38) Jaffe, A.; Ziebel, M. E.; Halat, D. M.; Biggins, N.; Murphy, R. Biomass. Chemosphere 2004, 57, 1225−1229.
A.; Chakarawet, K.; Reimer, J. A.; Long, J. R. Selective, High- (56) Esrafili, L.; Safarifard, V.; Tahmasebi, E.; Esrafili, M. D.;
Morsali, A. Functional Group Effect of Isoreticular Metal−Organic
Temperature O2 Adsorption in Chemically Reduced, Redox-Active
Frameworks on Heavy Metal Ion Adsorption. New J. Chem. 2018, 42,
Iron-Pyrazolate Metal-Organic Frameworks. J. Am. Chem. Soc. 2020,
8864−8873.
142, 14627−14637. (57) Safari, M.; Yamini, Y.; Masoomi, M. Y.; Morsali, A.; Mani-
(39) Ghanbari, T.; Abnisa, F.; Wan Daud, W. M. A. A Review on Varnosfaderani, A. Magnetic Metal-Organic Frameworks for the
Production of Metal Organic Frameworks (MOF) for CO2 Extraction of Trace Amounts of Heavy Metal Ions Prior to Their
Adsorption. Sci. Total Environ. 2020, 707, 135090. Determination by ICP-AES. Microchim. Acta 2017, 184, 1555−1564.
(40) Xu, G.-R.; An, Z.-H.; Xu, K.; Liu, Q.; Das, R.; Zhao, H.-L.
Metal Organic Framework (MOF)-Based Micro/Nanoscaled Materi-
als for Heavy Metal Ions Removal: The Cutting-Edge Study on
Designs, Synthesis, and Applications. Coord. Chem. Rev. 2021, 427,
213554.
(41) Grape, E. S.; Flores, J. G.; Hidalgo, T.; Martínez-Ahumada, E.;
Gutiérrez-Alejandre, A.; Hautier, A.; Williams, D. R.; O’Keeffe, M.;
Ö hrström, L.; Willhammar, T.; Horcajada, P.; Ibarra, I. A.; Inge, A. K.
A Robust and Biocompatible Bismuth Ellagate MOF Synthesized
under Green Ambient Conditions. J. Am. Chem. Soc. 2020, 142,
16795−16804.
(42) Marchenko, Z. Photometric Determination of Elements; MIR,
1971.
(43) Ayawei, N.; Ebelegi, A. N.; Wankasi, D. Modelling and
Interpretation of Adsorption Isotherms. J. Chem. 2017, 2017,
3039817.
(44) Simonin, J.-P. On the Comparison of Pseudo-First Order and
Pseudo-Second Order Rate Laws in the Modeling of Adsorption
Kinetics. Chem. Eng. J. 2016, 300, 254−263.
(45) Namdeo, M.; Bajpai, S. K. Chitosan-Magnetite Nano-
composites (CMNs) as Magnetic Carrier Particles for Removal of
Fe(III) from Aqueous Solutions. Colloids Surf., A 2008, 320, 161−
168.

9778 https://doi.org/10.1021/acssuschemeng.1c02146
ACS Sustainable Chem. Eng. 2021, 9, 9770−9778

You might also like