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

Dalton Transactions

Two-Dimensional MAX-Derived Titanate Nanostructures for


Efficient Removal of Pb(II)

Journal: Dalton Transactions

Manuscript ID DT-ART-10-2018-004301.R1

Article Type: Paper

Date Submitted by the


19-Dec-2018
Author:

Complete List of Authors: Wang, Xiangke; North China Electric Power University, College of
Environment and Chemical Engineering; King Abdulaziz University, NAAM
Research Group, Faculty of Science
Gu, Pengcheng; Chinese Academy of Sciences,
Zhang, Sai; North China Electric Power University
Zhang, Chenlu; a. College of Environmental Science and Engineering,
North China Electric Power University
Wang, Xiangxue; School of Environment and Chemical Engineering,
North China Electric Power University,
Khan, Ayub; North China Electric Power University, College of
Environmental Science and Engineering
Wen, Tao; North China Electric Power University, School of Environment
and Chemical Engineering
Hu, Baowei; Shaoxing University,
Alsaedi, Ahmed; NAAM Research Group, Faculty of Science, King
Abdulaziz University,
Hayat, Tasawar; e. NAAM Research Group, Faculty of Science, King
Abdulaziz University, Jeddah 21589, Saudi Arabia
Please do not adjust margins 
Page 1 of 8 Dalton Transactions

Journal Name   

ARTICLE 

Two‐Dimensional MAX‐Derived Titanate Nanostructures for 
Efficient Removal of Pb(II) 
TReceived 00th January 20xx, 
Accepted 00th January 20xx 
Pengcheng Gu,a,b Sai Zhang,a Chenlu Zhang,a Xiangxue Wang,a Ayub Khan,a  Tao Wen,a* Baowei 
Hu,b* Ahmed Alsaedi,c Tasawar Hayat,c and Xiangke Wanga,b* 
DOI: 10.1039/x0xx00000x 
Two‐dimensional (2D) nanomaterials have been well identified as one of the promising star materials benifting from  their 
www.rsc.org/  great promise for waste treatment. Nowadays, the investigation of wastewater remediation is highly imperative and still 
challenging. Here, a novel class of 2D MAX@titanate nanocomposits was fabricated by a simple oxidation and alkalization 
method,  which  exhibited  different  morphologies  and  impressive  elimination  performance.  The  Pb(II)  uptake  processes 
were  dramatically  affected  by  solution  pH  and  reached  equilibrium  quickly.  Abundant  functional  groups  and  enhanced  
specific surface areas endowed T‐NTO nanofibers with an outstanding adsorption capacity of 328.9 mg g‐1 at pH= 5.0 and 
T= 298K, which was much higher than that of T‐KTO nanoribbons. Moreover, the possible mechanism was expounded with 
aid of Raman, FT‐IR, XRD and XPS analyses, in which the synergistic effect of surface complexation and ion exchange made 
tremendous contribution to the adsorption performation. On the basis of above analyses, this work not only presented a 
novel and facile strategy for preparing T‐NTO and T‐KTO nanostructures as superior adsorption capacity, but aslo broaden 
other functional MAX‐derived nanostructures’prospective application in environmental cleanup. 

greatly hindered its inherent capacity in the purification of


Introduction  wastewater.15  To address this issue, numerous investigations have
The contamination of heavy metal ion is a significant public and been devoted to design functionalized Ti3AlC2 (TAC) derived
environmental concern due to its underlying threaten to human nanostructure for optimizing its adsorption performances, for
beings and the ecological environment.1-3 Among the heavy metal instance, the synthesis of hydroxylated Ti3C2Tx by an alkalization-
pollutants, Pb(II) ions, which are highly toxic and capable of intercalation strategy and fabricating hydrated Ti3C2Tx as a high
bioaccumulation, could cause serious ecological and health issues efficient U(VI) adsorbent.16,17 Although these optimization methods
even in trace concentration.4-6 Hence, remediation of wastewater were skillful and effective for increasing the functional groups of
containing Pb(II) manifested a great practical value for mitigating Ti3C2Tx, the fluoride-containing etching procedure inevitably limited
the risks to the human beings and environmental ecosystem.7,8 its practical applications.18 Therefore, designing novel functionalized
Currently, two-dimensional (2D) materials have prompted great TAC-derived materials with well layered nanostructures using a
research interests by virtue of their unique and impressive structural straightforward and eco-friendly method, was of great importance
properties.9,10 Specifically, MAX-phase-derived 2D materials, for wastewater remediation. Owing to the fact that the unique
referred as “Mxenes”, acted a vital role in the optimization of structures and abundant oxygen-containing groups, titanate
rechargeable batteries and supercapacitors, etc.11,12 As the promising nanomaterial were regareded as the promising ones for capturing
star of energy storage, Ti3C2Tx (Tx = -F, -OH) was most studied pollutants.19,20 Hence, the combination of TAC with abundant active
subset of “Mxene” materials, which was exfoliated by etching the sites of titanate materials was a wise and effective choice. To our
“Al” element from its Ti3AlC2 precursor.13,14 Unfortunately, the best knowledge, few studies regarding the elimination of heavy
absence of abundant functional groups, introduction of fluoride ions, metal ions with TAC-derived titanate naonmaterials have been
published.
Inspired by this point, herein, two types of TAC@titanate
a. MOE Key Lab of Resources and Environmental System optimization, College of  composites were successfully fabricated by the simultaneous
Environmental Science and Engineering, North China Electric Power University,  alkalization and oxidation of irregular TAC directly. The synthetic
Beijing 102206, P.R. China 
b. School of Life Science, Shaoxing University, Huancheng West Road 508, Shaoxing  process of TAC@titanate composites was shown in Fig. 1. The
312000, P.R. China  introduction of multi-functional oxygen-containing groups on TAC
c. NAAM Research Group, Faculty of Science, King Abdulaziz University, Jeddah 
by fluoride-free strategy was undoubtedly beneficial for the
21589, Saudi Arabia NAAM Research Group, Faculty of Science, King Abdulaziz 
University, Jeddah 21589, Saudi Arabia  enhancement of adsorption performance.21 Systematic investigations
*Corresponding authors.  of adsorption process of Pb(II) on the as-prepared T-NTO and T-
E‐mail:  xkwang@ncepu.edu.cn  (X.K.  Wang);  twen@ncepu.edu.cn  (T.  Wen);
hbw@usx.edu.cn (B. Hu). 
KTO were evaluated by batch experiments, which involved the
Tel(Fax): 86‐10‐61772890.  effect of pH, contact time and temperature, etc. Moreover, the
See DOI: 10.1039/x0xx00000x 

This journal is © The Royal Society of Chemistry 20xx  J. Name., 2013, 00, 1‐3 | 1 

Please do not adjust margins 
Please do not adjust margins 
Dalton Transactions Page 2 of 8

ARTICLE  Journal Name 

adsorption mechanisms were subsequently verified by Raman, FT-  


IR and XPS analyses. The investigation herein endowed the TAC-
derived nanostructure a promising material for management of Pb(II)
from aqueous solution. Result discussion  
Characterization 
NaOH&H2O2 KOH&H2O2 The structures and surface morphologies of the pristine Ti3AlC2, T-
NTO and T-KTO samples were characterized by TEM and SEM. As
evidenced from Fig. 2, the SEM image of Ti3AlC2 showed typical
T-NTO TAC T-KTO
bulk irregular shapes (Fig. 2b). The surface morphology was
dramatically transformed after oxidation and alkalization
hydrothermal treatment. Remarkably, by varying the alkaline
Fig. 1 Schematic diagram of T‐NTO and T‐KTO nanostructure formation.   medium (NaOH for T-NTO and KOH for T-KTO),  the products
exhibited different morphologies of urchin-like nanofibers (Fig. S1a,
Fig. S1b and Fig. 2d) and kelp-like nanoribbons (Fig. S1c and Fig.
Experimental section 
2g), respectively. In addition, a high-magnification SEM image
Preparation of adsorbents  further revealed the morphology of T-NTO nanofibers and T-KTO
Typocally, a mixture of 200 mg of Ti3AlC2 powders (purchased nanoribbons (Fig. 2a and Fig. 2c). High-resolution TEM (HRTEM)
from Forsman Technology Company), 75 ml of 5 M (12 g) NaOH, images of margin zone (Fig. 2e) further manifested the lattice
and 3ml of 30% H2O2 was placed into a beaker. After magnetic spacings of 0.75 nm, corresponding to the (001) planes of the NTO
stirring for 8 h, the suspension was transfeered to a Teflon-lined phase.22 Meanwhile, an interlayer spacing of 0.86 nm, corresponding
stainless-steel reaction vessel, and then be kept in an oven a to the (200) plane of KTO, was presented in Fig 3i.23 However, the
temperature of 180 oC for 24 h. The gray product was collected by sheet of intermediate zone is too thick to pierce to the attacking of
deionized water until the pH decreased to 6. Likewise, T-KTO was the electron. The TEM image of Fig. S1d showed that the surface of
prepared by an analogous procedure using KOH (15 g) as alkali TAC was very smooth. Based on the above characterization, a
source. The detailed information for the characterization was possible conclusion could be that NTO and KTO phases occurred on
described in SI. the magin of TAC after oxidation and alkalization treatment, which
was attributed to the super high chemical stability of TAC.24
Adsorption experiments 
Moreover,  the color also changed from black to gray and light
A typical batch experiment was as follows, the target Pb(II) stock yellow, suggesting the successful transformation to T-NTO and T-
solution (300 mg/L), T-NTO or T-KTO suspension (1.2 g/L), KTO.
deionized water and background ions (NaNO3) were successively
added to a 10mL polyethylene tube to achieve the desired
concentrations. The suspensions were shaken for 24 h and separated
by filtration. The Pb(II) concentration was detected by a
spectrophotometric method at the wavelength of 616 nm. The
solution pH was adjusted by the addition of 0.005 M-0.1 M HNO3 or
NaOH. For the adsorption isotherm studies, the adsorption
experiments were explored under different initial concentrations (5-
70 mg/L) at pH = 5.0. Kinetics investigation of T-NTO and T-KTO
(m/V= 0.2 g/L) were determined at various contact time (2-360
min). Details for the corresponding kinetic models employed in this
study were listed in SI. The adsorption percentage (%) and the
adsorption amount (qe, mg g-1) were caculated from the following
Fig. 2 (a‐c) SEM images of T‐NTO Ti3AlC2, and T‐KTO; (d,g) TEM images of T‐
equations: NTO and T‐KTO; (e, f) HRTEM images of T‐NTO and T‐KTO.
V   C0  Ce  Furthermore, the crystal structure transformation was also
qe  (1) examined by XRD (Fig. 3). Very sharp diffraction peaks at 9.52°,
m
19.2°, 33.9°, and 39.0° were assigned to the Ti3AlC2 crystal phase

Adsorption% 
 C0  Ce  100% (2)
(PDF card 52-0875), well consistent with previous reports. 25,26 It
was found that these typical peak intensities of Ti3AlC2 were
C0
significantly decreased after oxidation and alkalization treatment. 
Meanwhile, new peaks at 2θ = 9.87°, 24.36°, 28.38° and 34.63° were
where C0 (mgꞏL-1) is the preliminary concentration and Ce (mgꞏL-1)
indexed to (100), (201), (111) and (112) plane of typical primitive
is the equilibrium concentration after adsorption, m (mg) represents
monoclinic Na2Ti3O7 (NTO) (PDF card 31-1329), respectively.27-29
the mass of T-NTO or T-KTO, V (mL) is the volume of the
Similarly, the appearance of peaks at 10.07°, 23.97°, 28.96° and
suspension.
47.96° corresponded to (200), (110), (402), and (020) plane of
monoclinic K2Ti4O9 (KTO) phase (PDF Number 32-0861).30 Despite

2 | J. Name., 2012, 00, 1‐3  This journal is © The Royal Society of Chemistry 20xx 

Please do not adjust margins 
Please do not adjust margins 
Page 3 of 8 Dalton Transactions

Journal Name  ARTICLE 

the formation of main peaks, it can be readily observed that some of and morphology.24 As calculated by pore structure parameters, the
the characteristic reflections of NTO and KTO were still missing. average pore sizes were situated at diameters of 3.7 nm for T-NTO
The missing reflections were probably due to low-concentration and 3.5 nm for T-KTO, indicting the existence of mesopores.37
impurity phases and partial crystallization of titanate phase, as also Benefiting from the facile preparation method, abundant functional
found by previous results.31-32 To determine the components of T- groups and increased specific surface areas, such MAX-derived
NTO and T-KTO, the obtained nanocomposites were dissolved into titanates were highly expected to be the promising candidates for
20 M HNO3 solution, respectively. The molar ratio of metal elements wastewater remediation.  
was measured by inductively coupled plasma optic emission
spectrometer (ICP-OES). Thus, Na2Ti3O7 component in T-NTO was
~ 49.2 wt%, while content of K2Ti4O9 in T-KTO was ~ 44.4 wt%
based on ICP analysis (Tables S1). The valence state of as-prepared
materials was further determined with the aid of XPS investigations.
As shown in Fig. 4a, an apparent Na 1s signal at 1017.0 eV and K 2p
signal (K 2p1/2 ∼295.1 eV, K 2p3/2 ∼292.3 eV) were emerged for T-
NTO and T-KTO, respectively.27 The peaks at approximately 463.8
eV and 458.0 eV corresponding to Ti 2p1/2 and Ti 2p3/2 could be
ascribed to the formation of oxidized Ti(IV) (Fig. 4b).33 The Raman
spectra (Fig.4c) were in accordance with those of titanate materials
with preious report. The peaks located at 885 cm−1 and 923 cm−1 in  
T-NTO were triggered by short Ti-O stretching vibrations
Fig 3. XRD patterns of Ti3AlC2, T‐NTO and T‐KTO. 
coordinated with Na+.34 Similarly, for the T-KTO, the peak at 903
cm−1 was the characteristic of the Ti-O stretching vibration related to Effect of solid content  
K+.35 The peak at 442 cm-1 represented the Ti-O bending vibration
involving six-coordinated titanium and three-coordinated oxygen It is well-known that an appropriate solid content (m/V) was a
atoms in the [TiO6] octahedron. FT-IR spectra were employed for the crucial factor for influencing the adsorption process.38 As graphed in
chemical constitutions on the T-NTO and T-KTO. Peaks at 900 cm-1 Fig. 5a, the adsorption percentage of Pb(II) underwent a remarkable
and 953 cm-1 for both T-NTO and T-KTO were typical the stretching increase from ~ 71.8% to ~ 95.8% with T-NTO content increased
vibrations of Ti-O involving non-bridging oxygen atoms.32 The from 0.05 to 0.8 g/L. The adsorption behavior of Pb(II) on T-KTO
bands at ~1396 cm-1 and 3410 cm-1 were characteristic of the presented an analogous tendency to that of Pb(II) on T-NTO (Fig.
bending vibrations of Ti-OH and stretching vibration of -OH, 5b). Since the adsorption efficiency of T-NTO can reach 92.3% at
respectively (Fig. 4d).36 0.2 g∙L-1, which is seen as an appropriate solid content for the
The foregoing analyses confirmed the successful fabrication of T- following batch investigations. The batch experiments of T-KTO are
NTO and T-KTO during the simultaneous oxidation and alkalization also carried out under this condition for the fair comparison. It was
treatment. To explore the proposed mechanism for the titanate necessary to point out that the Qe values of Pb(II) on T-NTO and T-
formation process, the XRD patterns were monitored at different KTO decreased obviously with the increase of m/V. At high m/V
reaction condition and shown in Fig. S2. As for the oxidized Ti3AlC2 conditions, the available adsorption sites of adsorbents would be
without involving NaOH treatment (denoted as H2O2-TAC), an constrained, which might be responsible for the decreased Qe value.
appearance of TiO2 peak was successfully recorded as comparison Another reason may be the aggregation of solid particles, further
with the pristine Ti3AlC2. However, no significant change was reducing the complexation action between the binding sites and
observed on the alkalized Ti3AlC2 (denoted as NaOH-TAC). This Pb(II).39
phenomenon verified that the surface of TAC was partially oxidized
with aid of H2O2. Herein, a reasonable mechanism was proposed for
the formation process of TAC-derived NTO and KTO. As the
precursor, The surface of TAC could be partially converted into
titanium oxycarbides (TiO2) during the rapid oxidation treatment.
And then, the sodium/potassium titanates were further formed under
the alkalization process under the alkalization process of titanium
oxycarbides. The transformation of chemical mechanisms might be
concluded as the following equations :
Ti3AlC2 + x H2O2 = Ti3- x AlC2 + xTiO2 + x H2↑ (a)
3TiO2+2NaOH = Na2Ti3O7+H2O (b1)
4TiO2+2KOH = K2Ti4O9+H2O (b2)
Subsequently, the N2 adsorption-desorption isotherms were shown in
the Fig. 4e and 4f, the BET specific surface areas of T-NTO and T-
KTO were determined to be 62.4 m2/g and 57.8 m2/g, respectively. It
was worth mentioning that these values were superior to those of
pristine TAC, suggestive of the successful reformation of structure

This journal is © The Royal Society of Chemistry 20xx  J. Name., 2013, 00, 1‐3 | 3 

Please do not adjust margins 
Please do not adjust margins 
Dalton Transactions Page 4 of 8

ARTICLE  Journal Name 

Fig. 5 (a, b) Effect of solid content on Pb(II) adsorption to T‐NTO and T‐KTO; (c, 
a O 1s
b
Ti 2p3/2
d)  The  kinetics  studies  of  Pb(II)  adsorption  on  T‐NTO  and  T‐KTO.  All 
Ti 2p
T-KTO T-KTO Ti 2p1/2 experiments were performed at T = 298K, pH = 5.0 ± 0.1, C0(Pb(II)) = 10 mg/L, 
Intensity (a.u.) K2p
C(NaNO3) = 0.01 M. 

Intensity (a.u.)
Na 1s

T-NTO T-NTO
Effect of pH  
Generally speaking, solution pH as an important factor affect the
0 400 800
Binding energy (eV)
1200 450 460
Binding energy (eV)
470 surface charge of adsorbents, existent forms of Pb(II) species and
c T-NTO d 442
T-NTO 886 917
protonation degree of surface functional groups.40 The effect of pH
900
1396 T-NTO+Pb2+ 3460 towards Pb(II) on T-NTO and T-KTO was carefully investigated. As
T-NTO+Pb2+
Intensity (a.u.)

Intensity (a.u.)
907
T-KTO observed in Fig. 6a, the uptake percentage of T-NTO increased
442
rapidly from ∼44% to ∼92% with increasing pH (2.5-6.5),
953 T-KTO 903
1396 T-KTO+Pb2+

960
T-KTO+Pb2+ 870
maintained the elevated level, followed by a slight descend at pH >
1000 2000 3000 4000 400 800 9.5. Similar tendency was also observed for the T-KTO/Pb systems.
Wavelength (cm-1) Raman shift (cm-1)
200
e
120
f The above results can be explained by the inherent properties of the
Quantity Adsorption (cm /g SPT)

Adsorption Adsorption
Quantity Adsorption (cm3/g SPT)

Desorption
150
Desorption
adsorbents and the species distribution of Pb(II) in  aqueous solution.
3

V/dlog (D) Pore Volume (cm /g)

0.04
V/dlog (D) Pore Volume (cm /g)

0.02
3

80
3

3.7 nm 3.5 nm

100
0.01 0.02 Obviously, the point of zero charge (pHpzc) of T-NTO and T-KTO
50
0.00

0 40 80
Pore Diameter (nm)
120
40
0.00

0 40 80
Pore Diameter (nm)
120
were approximately 2.75 and 3.02, indicative of both the two
62.4 m2/g 57.8 m2/g
products carried the highly positive surface charge under strong
0
0.0 0.2 0.4 0.6 0.8 1.0
0
0.0 0.4 0.8 acidic condition (Fig. 6b). Calculated by Visual MINTEQ ver. 3.0,
Relative Presure (P/P0) Relative Presure (P/P0)
the relative distribution of Pb(II) species were shown in the Fig. S3.
It was noteworthy that the positive charged (Pb2+) was the major
Fig. 4 (a) XPS spectra of T‐NTO and T‐KTO; (b) Ti 2p of T‐NTO and T‐KTO; FT‐ existing species at pH < 6.5. Accordingly, the electrostatic repulsion
IR spectra (c) Raman spectra (d) of T‐NTO and T‐KTO before and after Pb(II) 
removal;  (e,  f)  N2  adsorption‐desorption  isotherms  of  T‐NTO  and  T‐KTO 
together with competition with H+ were mainly responsible for the
(insets are the pore size distribution).  unsatisfactory removal performance at low pH. As pH increased
from pHPZC to 6.5, the enlarged adsorption percentage probably
Kinetics studies 
arose from the effective electrostatic attraction between negative
Fig. 5c and 5d diagrammed the plots of adsorbed Pb(II) versus the adsorbents surface and positive charged Pb(II) species (Pb2+ and
adsorption time. It was worth noting that the removal percentages of Pb(OH)+).41-43 While at high pH values, several negatively charged
Pb(II) on T-NTO and T-KTO reached 90.0% and 68.0% within the OH- were capable of competing with negatively charged Pb(OH)3-,
initial 1h, then the equilibrium state being achieved slowly. The high leading to the negative effect on ion exchange, which further
elimination rates of Pb(II) in the initial stage were mainly associated impeded the elimination of Pb(II) process. 5,44
with the abundant free available adsorption sites on the surface. As 10
summarized in Table S2, the adsorption rate of T-NTO towards 100 a T-NTO b T-NTO
T-KTO 0 T-KTO
Pb(II) was superior as compared with that of T-KTO, which might
Zeta potential (mV)
Adsorption %

80
be attributed to the former’s highly active sites, further boosted the -10

adsorption energy between adsorbents and Pb(II). To get insight into -20
60
the underlying adsorption behavior, the kinetic data was mimicked -30
by two typical kinetic models, which were exhibited in SI section.
40 -40
The plot of t/Qt vs t (h) disclosed a perfect linear relation (correlation 2 4 6
pH
8 10 2 4 6
pH
8 10
400
coefficient values R2 = 0.999), which was a typical characteristic for c d
300
pseudo-second-order processes (insets of Fig. 5c and 5d), suggesting 300

the adsorption process was substantially under a chemical reaction.


Qe (mg/g)
Qe (mg/g)

200
200
100 298 K 298 K
100 a 300 b 160 313 K 313 K
90 100 328 K 100 328 K
Langmuir Langmuir
80 Freundlich Freundlich
120
Adsorption (%)
Adsorption (%)

90 200
Qe (mg/g)

0
Qe (mg/g)

70 0
0 15 30 45
80 0 15 30 45
60 Ce (mg/L)
80 Ce (mg/L)
100
50
40
Fig.  6  (a)  Effect  of  pH  on  Pb(II)  adsorption  to  T‐NTO  and  T‐KTO;  (b)  Zeta 
70 40
0
0.0 0.3 0.6 0.9
potential of T‐NTO and T‐KTO; (c, d) Adsorption isotherms of Pb(II) on T‐NTO 
0.0 0.3 0.6 0.9
Solid content (g L-1) Solid content (g L-1)
and  T‐KTO  calculated  at  298K,  313K  and  328K.  Conditions:  pH  =  5.0  ±  0.1, 
100
m/V = 0.2 g/L, C(NaNO3) = 0.01 M. 
c 80
d
80
Thermodynamic study 
Adsorption (%)
Adsorption (%)

0.12
0.12 60
60
With the aim of exploring adsorption mechanisms and evaluating
t/Qt (h g/mg)

t/Qt (h g/mg)

0.08 0.08

40 0.04 40 0.04 adsorption performance, the adsorption thermodynamic studies were


20
0.00
0 4 8 12 20
0.00
0 4 8 12
exhibited in Fig. 6c and 6d. It was quite clear that the higher
T (h)

0 4 8 12 0 4 Time (h) 8
T (h)

12
temperature was beneficial to the adsorption performance of Pb(II)
Time (h)
(328K > 313K > 298K). the Langmuir and Freundlich models were

4 | J. Name., 2012, 00, 1‐3  This journal is © The Royal Society of Chemistry 20xx 

Please do not adjust margins 
Please do not adjust margins 
Page 5 of 8 Dalton Transactions

Journal Name  ARTICLE 

used for fitting the adsorption data and the related equations were To probe the adsorption mechanism specifically, FT-IR and Raman
exhited as following : 45,46 spectra were collected on T-NTO and T-KTO before and after
adsorption (Fig. 4). The Raman spectra after Pb(II) adsorption
Langmuir model: qe 
K LQmax Ce (3) (labelled as T-NTO + Pb2+ and T-KTO + Pb2+) clearly illustrated a
1  K LCe remarkable change of molecular structure information (Fig. 4d). The
peaks at 886 cm-1 and 917 cm-1 (Ti-O-Na) merged and shifted to
Freundlich model: Qe  K f Ce1 n (4) lower intensity, providing a direct proof for the fractional
replacement of Na+ by Pb(II).54,55 Similarly, the changed Ti-OK peak
where Ce (mg/L) is the Pb(II) concentration at equilibrium, qmax at 903 cm-1 for T-KTO was attributed to the short Ti-O vibrations
(mg/g) represents the topmost adsorption capacity on theoretical affected by the target ions. However, there are no obvious change or
level, and qe (mg/g) represents the amount of Pb(II) adsorbed on the shift towards the peaks of 442 cm-1 after adsorption, signifying the
adsorbents. b (L/ mg) and n are considered as the constants of the stability of [TiO6] octahedron in the adsorption process. Those
Langmuir and Freundlich models, respectively. The related phenomena confirmed that ion exchange between Pb(II) and the
parameters were summarized in Table 1, all manifested a well interlayered Na+ ions was the key mechanism for Pb(II) removal.56
agreement with the Langmuir model, indicating that immobilization As for the FT-IR spectra (Fig. 4c), upon adsorption of Pb (II), it was
of Pb(II) process was a monolayer coverage process.47 The topmost clear that the peaks of -ONa (900 cm-1) and -OK (953 cm-1) groups
capacity of Pb(II) adsorption on T-NTO was calculated to be 328.9 shifted to 907 cm-1 and 960 cm-1, due to ion exchange of Na+ /K+ in -
mg g-1, which was larger than that of T-KTO (248.3 mg g-1). One ONa/OK with Pb (II).56 The XRD patterns after Pb(II) uptake were
possible explanation, which is proposed for the T-NTO with superior shown in Fig. S4, the interlayer spacing (d100 spacing) of T-NTO
adsorption performance, was that the larger specific surface area decreased from 0.902 nm to 0.864 nm caused by ion exchange
could offer more binding locations to promote Pb(II) uptake.48 An strongly, suggesting that the replacement of Na+ in the interlayer of
alternative explanation was considered from the variance of radius T-NTO resulted in the interlayer shrinking.57 Similarly, the d spacing
between K+ and Na+ in the titanate surface and interlayer. It was of T-KTO decreased from 0.884 nm to 0.830 nm. To better clarify
well known that the ion-exchange is more favorable between the the cause of the dissimilarities of elimination capacity on T-NTO
adsorbate and ions with smaller radius size.4,29 49 Hence, Na+ with and T-KTO deeply, XPS spectra were further recorded in Fig. 7a and
smaller radius (0.102 nm), was more prone to complete the ion 7b, which were calibrated by C 1s peak at 284.6 eV (see Fig. S5).
exchange with Pb(II), thereby endowing T-NTO with a higher As seen from Fig. 7a, the decrement of Na 1s (from 16.53% to
adsorption efficiency for Pb(II) than T-KTO, which was also 10.31%) and K 2p (from 12.54% to 8.76%) peaks, together with the
reported by Xu et al..50 In addition, as compared with majority of appearance of two peaks of Pb 4f, which further reasonably
previously reported nanomaterials, the extraordinary adsorption confirmed the ion-exchange mechanism, being in good agreement
capacity of T-NTO makes it one of the most effective Pb(II) with Raman, FT-IR and XRD results.19 Similar behavior was also
adsorbents. 51,52 reported by Liu et al.58,59 and other references.60-63 In the high-
Herein, the thermodynamic parameters, i.e., ∆G0 (free energy), resolution of O 1s spectra (Fig.7c and 7d), the peak of 529.6 eV was
∆H0 (enthalpy change) and ∆S0 (entropy change), were calculated to corresponded to the crystal lattice [TiO6]. Additionally, the peak
reveal the thermodynamic viability (see the detailed information in located at 530.1 eV was identified to Ti-OH, which can be assigned
the SI). As summarized in Table S3, the positive ΔH0 values further to the surface complexation site.64 After Pb(II) adsorption, the
proved that the elimination process of Pb(II) on T-NTO and T-KTO binding energy of Ti-OH (530.1 eV ) was shifted to higher binding
was an endothermic reaction. Meanwhile, it could be well accepted energy for T-NTO (530.4 eV) and T-KTO (530.2 eV), respectively,
that the adsorption reaction was spontaneous, as a result of the which could be traceable to the donation of electrons from O 1s and
positive ∆S0 and negative ΔG0 values.53 Attractively, the ∆G0 values the formation of Pb-O bonding. The finding herein suggested that
on T-NTO were more negative than those of Pb(II) on T-KTO, the surface complexation site plays important role during the
reflecting that the adsorption process of Pb(II) on T-NTO was more adsorption process of Pb(II).65 As summarized in Table 2, it was
spontaneous as comparison with T-KTO. evident that the Ti 2p peaks of T-NTO and T-KTO also moved to
Table 1 The Parameters for the thermodynamic models of Pb(II) adsorption  higher binding energies, which were caused by the donation effect of
on T‐NTO and T‐KTO.  the outmost O and the subsequent induce effect between Ti-OH (Fig.
  7e and 7f). Notably, the Ti 2p bonding shift of T-NTO (0.4 eV) was
T-NTO T-KTO larger than that of T-KTO (0.2 eV), which was in good agreement of
O 1s, indicating a stronger interaction between T-NTO and Pb(II).
Parameters 298K 313K 328K 298K 313K 328K
In brief, T-NTO possessed higher adsorption capacity than T-KTO
Qe,max owing to the former’s higher affinity and stronger complexation
Langmuir 246.6 302.7 341.8 305.1 340.8 394.8
(mg∙g-1)
R 2
0.981 0.955 0.996 0.975 0.975 0.956
force to bind with Pb(II). These finding was well in accordance with
the thermodynamics results.
Freundlich 1/n 0.36 0.39 0.48 0.47 0.47 0.52
2
R 0.823 0.812 0.935 0.892 0.905 0.915

Interaction mechanism  

This journal is © The Royal Society of Chemistry 20xx  J. Name., 2013, 00, 1‐3 | 5 

Please do not adjust margins 
Please do not adjust margins 
Dalton Transactions Page 6 of 8

ARTICLE  Journal Name 

Raman and XPS), the possible mechanism was well elaborated, in


a Pb 4f T-NTO+Pb(II)
b Pb 4f7/2 Pb 4f5/2 which the synergistic effect of surface complexation and ion
exchange made tremendous contribution to the excellent adsorption
Intensity (a.u.)

T-NTO T-NTO+Pb(II)

Intensity (a.u.)
performation. More importantly, benefiting from the higher specific
Pb 4f T-KTO+Pb(II)
Pb 4f7/2
Pb 4f5/2 surface area, stronger complexation capacity and smaller size of
T-KTO
T-KTO+Pb(II) interlayer ions, the resulting T-NTO nanofibers exhibited a superior
130 140 150
potential for Pb(II) adsorption. Therefore, the multiple advantages,
0 400 800 1200
Binding erergy (eV) Binding energy (eV) including fluoride-free methodology and high adsorption capacities,
c TiO6 Ti-OH
O 1S d TiO6 Ti-OH O 1S
made the MAX@titanate a promising candidate for environmental
remediation.

Intensity (a.u.)
T-NTO T-KTO
Intensity (a.u.)

T-NTO+Pb(II)
T-KTO+Pb(II)
Acknowledgements 
528 532 528 532
Binding energy (eV) Binding energy (eV) The financial support from the Science Challenge Project
e f Ti 2p (TZ2016004), the National Key Research and Development Program
Ti 2p T-KTO
T-NTO
of China (2017YFA0207002), NSFC (21607042, 21707033,
Intensity (a.u.)

21577032, 21777039), the Fundamental Research Funds for the


Central Universities (JB2017193, 2017MS045, 2017YQ001,
T-NTO+Pb(II) T-KTO+Pb(II)
2018ZD11), the NCEPU “Double First-Class” Graduate Talent
455 460 465
Cultivation Program” (035/XM1805316).
455 460 465
Binding energy (eV) Binding energy (eV)

Fig.  7  (a)  XPS  spectra  of  T‐NTO  and  T‐KTO  before  and  after  Pb(II)  removal;  References 
XPS spectra of Pb 4f (b), O 1s (c, d), and Ti 2p (e, f).  
  1 G. X. Zhao, J. X. Li, X. M. Ren, C. L. Chen and X. K. Wang, Environ Sci
The aforementioned results and analyses provided a strong proof Technol., 2011, 45, 10454-10462.
2 Y. D. Zou, X. X. Wang, Y. J. Ai, Y. H. Liu, Y. F. Ji, H. Q. Wang, T. Hayat, A
for the elimination of Pb(II) on T-NTO and T-KTO. As a result, the Alsaedi, W. P. Hu, X. K. Wang, J Mater Chem A, 2016, 4, 14170-14179.
ion exchange and surface complexation were the key adsorption 3 L. Y. Xiao, G. T. Shen, F. G. Mao, Y. W. Ling, C. Z. Jun, Y. C. Chang, G. S.
mechanisms for capturing Pb(II). Herein, taking the T-NTO as Wei, Acta Scientiae Circumstantiae, 2013, 25, 933-943.
4 Z. Chen, Y. Liang, D. S. Jia, W. Y. Chen, Z. M. Cui, X. K. Wang, Environ
example, the removal process of Pb(II) was expressed by the Sci: Nano, 2017, 4, 1851-1858.
following steps. Originally, the Pb(II) ions and the negatively 5 P. C. Gu, J. L. Xing, T. Wen, R. Zhang, J. Wang, G. X. Zhao, T. Hayat, Y. J.
charged surface sites are combined through the electrostatic Ai, Z. Lin, X. K. Wang. Environ Sci: Nano, 2018, 5, 946-955.
attraction. And then Pb(II) ions are assumed to exchange with Na+ in 6 G. X. Zhao, X. M. Ren, X. Gao, X. L. Tan, J. Li, C. L. Chen, Y. Huang, X.
K. Wang, Dalton Trans., 2011, 40, 10945-10952.
the layered titanate. The surface complexation between Pb(II) with 7 S. Song, S. Y. Huang, R. Zhang, Z. S. Chen, T. Wen, S. H. Wang, T. Hayat,
the surface functional groups (Ti-OH) of layer edges occurred in the A. Alsaedi, X. K. Wang, Chem Eng J., 2017, 325, 576-587.
adsorption process simultaneously. 8 Y. Z. Hu, X. X. Wang,Y. D. Zou, T. Wen, X. L. Wang, A. Alsaedi, T. Hayat,
X. K. Wang, Chem Eng J., 2017, 316, 419-428.
Table 2 Binding energies of elements in T‐NTO and T‐KTO before and after  9 P. C. Gu, S. Zhang, X. Li, X. X. Wang, T. Wen, R. Jehan, A. Alsaedi, T.
Pb(II) adsorption.  Hayat, X. K. Wang, Environ Pollut., 2018, 240, 493-505.
  10 S. Zhang, P.G. Gu, R. Ma, C.T. Luo, T. Wen, G.X. Zhao, W.C. Cheng, X.K.
Wang. Catal. Today, 2018. doi.org/10.1016/j.cattod.
O 1s Ti 2p 11 M. Boota, B. Anasori, C. Voigt, M. Q. Zhao, M. W. Barsoum, Y. Gogotsi,
Ti-OH [TiO6] Ti 2p1/2 Ti 2p3/2 Adv Mater., 2016, 28, 1517-1522.
T-NTO 530.1 eV 529.6 eV 463.6 eV 457.8 eV
12 X. F. Wang, S. Kajiyama, H. Iinuma, E. Hosono, S. Oro, I. Moriguchi, M.
Okubo, A. Yamada, Nat Commun., 2015, 6, 6544-6549.
T-NTO-Pb(II) 530.4 eV 529.7 eV 464.0 eV 458.2 eV 13 C. Eames, Islam MS, J Am Chem Soc., 2014, 136: 16270-16275.
T-KTO 530.1 eV 529.6 eV 463.8 eV 458.1 eV 14 M. R. Lukatskaya, O. Mashtalir, C. E. Ren, Y. Dall'Agnese, P. Rozier, P. L.
Taberna, M. Naguib, P. Simon, M. W. Barsoum, Y. Gogotsi, Science, 2013,
T-KTO-Pb(II) 530.2 eV 529.6 eV 464.0eV 458.3 eV
341, 1502-1505.
15 A. K. Fard, G. McKay, R. Chamoun, T. Rhadfi, H. Preud'Homme, M. A.
Atieh, Chem Eng J., 2017, 317, 331-342.
Conclusions  16 Q. M. Peng, J. X. Guo, Q. R. Zhang, J. Y. Xiang, B. Z. Liu, A. G. Zhou, R.
P. Liu, Y. J. Tian, J Am Chem Soc., 2014, 136, 4113-4116.
To conclude, presented above was a systematic investigation on the 17 L. Wang, W. Tao, L. Yuan, Z. Liu, Q. Huang, Z. Chai, J. K. Gibson, W. Q.
adsorption behavior of Pb(II) onto 2D MAX@titanate Shi, Chem Commun., 2017, 53, 12084-12087.
18 X. Wang, C. Garnero, G. Rochard, D. Magne,S. Morisset, S. Hurand, P.
nanostructures, which were synthesized by a straightforward
Chartier, J. Rousseau, T. Cabioc'h, C. Coutanceau, V. Mauchamp, S. Célérier,
oxidation and alkalization. Notably, the morphologies of TAC- J Mater Chem A, 2017, 5, 22012-22023.
derived titanate can be easily regulated by varying the type of alkali 19 L. Yin, P. Y. Wang, T. Wen, S. J. Yu, X. X. Wang, T. Hayat, A. Alsaedi, X.
source. The urchin-like T-NTO nanofibers and kelp-like T-KTO K. Wang, Environ Pollut., 2017, 226, 125-134.
20 D. J. Yang, Z. F. Zheng, H. Y. Zhu, H. W. Liu, X. P. Gao, Adv Mater., 2008,
nanosheets were successfully confirmed as effective adsorbents for 20, 2777-2781.
targeted removal of Pb(II), with a fast removal rate and outstanding 21 J. Li, Q. H. Fan, Y. J. Wu, X. X. Wang, C. L. Chen, Z. Y. Tang, X. K.
adsorption capacity. Based on the results of spectroscopic (FTIR, Wang, J Mater Chem A, 2016, 4, 1737-1746.

6 | J. Name., 2012, 00, 1‐3  This journal is © The Royal Society of Chemistry 20xx 

Please do not adjust margins 
Please do not adjust margins 
Page 7 of 8 Dalton Transactions

Journal Name  ARTICLE 

22 X. L Tan, M. Fang, L. Q Tan, H. N Liu, X. S. Ye, T. Hayat, X. K Wang, 60 A. Khan, J.L. Xing, A. M. Elseman, P.C. Gu, K. Gul, Y.J. Ai, R. Jehan, A.
Environ Sci: Nano, 2018, 5, 1140-1149 Alsaedi, T. Hayat, X.K. Wang. Dalton Trans., 2018, 47, 11327-11336.
23 Y. Liu, X Li, F. Li, Y. Z Liu, X. Y Yuan, L. F Zhang, S. W Guo, 61 X. Li, J.L. Xing, C.L. Zhang, B..Han, Y.H. Zhang, T. Wen, R. Leng, Z.H.
Electrochim. Acta, 2019, 295, 599-604 Jiang, Y.J. Ai, X.K. Wang. ACS Sustain. Chem. Eng., 2018, 6, 10606-10615.
24 X. H Xie, Y. Xue, L. Li, S. Chen, Y. Nie, W. Ding, Z. Wei, Nanoscale, 62 H.S. Zhu, X.L. Tan, L.Q. Tan, C.L. Chen, N.S. Alharbi, T. Hayat, M. Fang,
2014, 6, 11035-11040 X.K. Wang. ACS Appl. Nano Mater., 2018, 1, 2689-2698.
25 H. Lin, Y. W. Wang, S. S. Gao, Y. Chen, J. L. Shi, Adv Mater., 2018, 30, 63 H.S. Zhu, X.L. Tan, L.Q. Tan, H.F. Zhang, H.N. Liu, M. Fang, T. Hayat,
1703284-1703294, X.K. Wang. ACS Sustain. Chem. Eng., 2018, 6, 5206-5213.
26 A. Lipatov, M. Alhabeb, M. R. Lukatskaya, A. Boson, Y. Gogotsi, A. 64 H. Chen, X. X. Wang, L. X. Li, X. K. Wang, J Mater Chem A, 2013, 1,
Sinitskii, Adv Electro Mater., 2016, 2, 1600255-1600263.. 805-813.
27 G. D. Zou, J. X. Guo, X. Y. Liu, Q. R. Zhang, G. Huang, C. Fernandez, Q. 65 F. Liu, Y. J. Jin, H. B. Liao, L. Cai, M. P. Tong, Y. L. Hou, J Mater Chem
M. Peng, Adv Energy Mater., 2017, 7, 1700700-1700708. A, 2013, 1, 805-813.
28 S. Fu, J. f Ni, Y. Xu, Q. Zhang, L. Li, Nano Letters, 2016, 16, 4544-4551.
29 T. Wen, Z. W. Zhao, C. C Shen, J. Xing. Li, X. L Tan, A. Zeb, X. K Wang,
A. W. Xu, Scientific Reports., 2016, 6, 20920-20925
30 W. Q Cui, S. S Ma, L. Liu, J. S Hu, Y. H, Liang, J. G. McEvoy, Appl Surf
Sci., 2013, 271, 171-181
31 W. Wang, C. J Yu, Z. S Lin, J. G. Hou, H. M. Zhu, S, Q, Jiao, Nanoscale.,
2013, 5, 594-599.
32 L. Yin, S. Song, X. X. Wang, F. L. Niu, R. Ma, S. J. Yu, T. Wen, Y. T.
Chen, T. Hayat, A. Alsaedi, X. K. Wang, Environ Pollut.,2018, 238, 725-738.
33 D. O. Scanlon, C. W. Dunnill, J. Buckeridge, S. A. Shevlin, A. J. Logsdail,
S. M. Woodley, C. R. Catlow, Nat Mater., 2013, 12, 798-801.
34 V. B. Dmitry, M. Jens, A. L. Alexei, F. C. Walsh, Chem Mater., 2006, 18,
1124-1129.
35 Y. Dong, Z. S. Wu, S. Zheng, X. Wang, J. Qin, S. Wang, X. Shi, X. Bao,
ACS Nano, 2017, 11, 4792-4800.
36 M. Zhao, J. Huang, X. Guo, H. Chen, H. Zhao, L. Dong, X. J. Liu, J
Chem., 2015, 7, 1-12.
37 S. J. Yu, X. X. Wang, Z. S. Chen, J. Wang, S. H. Wang, T. Hayat, X. K.
Wang, J Hazard Mater., 2017, 321, 111-120.
38 P. C Gu, C. F. Zhao, T Wen, Y. J Ai, S. Zhang, W. Q Chen, J. Wang, B. W
Hu, X. K Wang, Chem. Eng. J. DOI: 10.1016/j.cej.2018.11.016
39 W. Yao, X. X. Wang, Y. Liang, S. J. Yu, P. C. Gu, Y. B. Sun, C. Xu, J.
Chen, T. Hayat, A. Alsaedi, X. K. Wang, Chem Eng J., 2018, 332, 775-786.
40 H. Chen, X. X. Wang, J. X. Li, X. K. Wang, J Mater Chem A, 2015, 3,
6073-6081.
41 K. R. Zhu, S. H. Lu, Y. Gao, R. Zhang, X.L. Tan, C. L. Chen, Appl Surf
Sci., 2017, 396, 1726-1735.
42 C. Y. Jing, J. L. Cui, Y. Y. Huang, A. G. Li, ACS Applied Mater Inter.,
2012, 4, 714-720.
43 G. X. Zhao, X. M. Ren, X. Gao, X. L. Tan, J. Li, C. L. Chen, Y. Huang, X.
K. Wang, Dalton Trans., 2011, 40, 10945-10952.
44 S. Y. Huang, S. Song, R. Zhang, T. Wen, X. C. Wang, S. J. Yu, W. Song, T.
Hayat, A. Alsaedi, X. K. Wang, ACS Sustain Chem Eng., 2017, 5, 11268-
11279.
45 X. Li, Y. Liu, C. Zhang, T. Wen, L. Zhuang, X. X. Wang, G. Song, D.
Chen, Y. J. Ai, T. Hayat, X. K. Wang, Chem Eng J., 2018, 336, 241-252.
46 R. Ma, L. Yin, L. Li, S. Zhang, T Wen, C. L Zhang, X. X Wang, Z. S Chen,
A. Alsaedi, X. K Wang, ACS Appl. Nano. Mater., 2018, 1, 5543-5552.
47 T. Wen, X. X. Wang, J. Wang, Z. S. Chen, J. Li, J. Hu, T. Hayat, A.
Alsaedi, B. Grambow, X. L. Wang, Inorg. Chem. Front., 2016, 3, 1227-1235.
48 Y. D. Zou, P. Y. Wang, W. Yao, X. X. Wang, Y. Liu, D. X. Yang, L. Wang,
J. Hou, A. Alsaedi, T. Hayat, X. K. Wang, Chem Eng J., 2017, 330, 573-584.
49 J. M. Gong, T Liu, X. Q.Wang, X. L Hu, L. Z. Zhang, Environ Sci
Technol., 2011, 45, 6181-6187.
50 J. S. Xu, H. Zhang, J. Zhang, E. J. Kim, J Alloy Compd., 2014, 614, 389-
393.
51 L. Xiong, C. Chen, Q. Chen, J. Ni, J Hazard Mater., 2011, 189, 741-748.
52 S. B. Yang, J. Hu, C. L. Chen, D. D. Shao, X. K. Wang, Environ Sci
Technol., 2011, 45, 3621-3627.
53 D. D. Shao, X. M. Ren, J. Wen, S. Hu, J. Xiong, T. Jiang, X. L. Wang, X.
K. Wang, J Hazard Mater, 2016, 302, 1-9.
54 M. Feng, W. You, Z. Wu, Q. Chen, H. Zhan, ACS Appl Mater Interf., 2013,
5, 12654-12662.
55 N. Li, L. D. Zhang, Y. Z. Chen, M. Fang, J. X. Zhang, H. M. Wang, Adv
Funct Mater., 2012, 22, 835-841.
56 R. Russo, D, Perry, Spetrochimica Acta, 1994, 50, 757-763.
57 D. J. Yang, Z. F. Zheng, H. W. Liu, H. Y. Zhu, X. B. Ke, Y. Xu, D. Wu, Y.
Sun, J Phys Chem C., 2008, 112, 16275-16280.
58 W. Liu, X. Zhao, T.Wang, D.Y. Zhao, J. R Ni. Chem Eng J., 2016, 286,
427-435.
59 W. Liu, X. Zhao, T.Wang, D.Y. Zhao, J, Fu, J. R Ni. J Mater Chem A,
2015, 3, 17676-17684.

This journal is © The Royal Society of Chemistry 20xx  J. Name., 2013, 00, 1‐3 | 7 

Please do not adjust margins 
Please do not adjust margins 
Dalton Transactions Page 8 of 8

ARTICLE  Journal Name 

Table of Contents 

Pb
Ion exchange
[≡Ti−O]− Na+ Pb2+
NaOH&H2O2

≡Ti−OH Pb2+
Surface complexation
TAC
T-NTO

8 | J. Name., 2012, 00, 1‐3  This journal is © The Royal Society of Chemistry 20xx 

Please do not adjust margins 

You might also like