You are on page 1of 15

Environmental Pollution 277 (2021) 116776

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Use of thermally modified waste concrete powder for removal of Pb


(II) from wastewater: Effects and mechanism*
Zihan Ma a, b, Runze Xue a, Jiang-shan Li b, c, *, Yaqin Zhao a, Qiang Xue b, c, Zhen Chen b,
Qiming Wang d, Chi Sun Poon c, d
a
School of Environment Science and Spatial Informatics, China University of Mining and Technology, No.1 Daxue Road, Xuzhou, 221116, Jiangsu, PR China
b
State Key Laboratory of Geomechanics and Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of Sciences, Wuhan, 430071,
China
c
IRSM-CAS/HK PolyU Joint Laboratory on Solid Waste Science, Wuhan, 430071, China
d
Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University, Hong Kong University, Hung Hom, Kowloon, Hong Kong,
China

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

Article history: Exploring effective uses of waste concrete powder (WCP), produced from recycling of construction &
Received 10 November 2020 demolition waste is beneficial to the environment and sustainable development. In this study, WCP was
Received in revised form first treated thermally to enhance the ability to remove Pb (II) from aqueous solutions. The experimental
11 February 2021
results revealed that the thermal treatment could enhance adsorption capacity due to modification of
Accepted 15 February 2021
Available online 18 February 2021
calcium bonding and pore structure of WCP. Preparation parameters such as temperature, particle size,
and water-cement ratio were investigated to obtain the optimal operational conditions. Batch adsorption
experiments were performed to explore influence factors of pH (1.00e6.00), ionic strength (0.05e2 mol/
Keywords:
Waste concrete powder (WCP)
L), dosage (2e50 g/L), and temperature (25e45  C). The pseudo-second-order kinetics model could
Lead adequately describe the adsorption process, and the Langmuir model was capable to predict the isotherm
Adsorption data well in the low concentration region (C0 < 500 mg/L). The maximum uptake capacity for Pb (II)
Modification calculated by Langmuir model at 25, 35 and 45  C were 46.02, 38.58 and 30.01 mg/g respectively, and the
removal rate of Pb (II) was 92.96% at a dosage of 50 g/L (C0 ¼ 1000 mg/L). Precipitation, ion exchange, and
surface complexation were identified to be the main mechanisms of Pb (II) adsorption through micro-
scopic investigation by SEM-EDX, XRD, FTIR, XPS, and BET inspections. The study confirms that the WCP
after thermal modification, can be selected as a promising adsorbent for the high performance and eco-
friendliness.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction bamboo, plastics, etc. Among these components, waste concrete


(WC) is the dominant one, including waste concrete powder (WCP)
With the rapid growth in economy and urbanization in many coming from the breakage of WC and recovery of coarse aggregates
countries, great volume of construction & demolition waste (CDW) (Kaliyavaradhan et al., 2020). The amount of WCP is increasing but
has been produced. The accumulation of construction waste causes recycling is rare due to the porous structure and high-water
disposal problems (Duan et al., 2019). About 1.8 billion tons of CDW adsorption of the waste. The waste concrete is usually randomly
are generated annually in China, and the amounts are also huge in dumped or disposed of, particularly in developing countries(Huang
other regions throughout the world (EPA, 2015; Xiao et al., 2018). et al., 2018), which brings about an urgent environmental burden.
CDW mainly consists of concrete, brick, rock, muck, mortar, For a long time, considerable literatures have demonstrated
some recycling potential of WCP as construction materials. With
notable reactivity, WCP can be reused as a supplementary cemen-
*
This paper has been recommended for acceptance by Jo €rg Rinklebe. titious materials (SCMs) to partially replace cement in limited
* Corresponding author. State Key Laboratory of Geomechanics and Geotechnical amount and retain the mechanical properties of the cementitious
Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of Sciences, products(Ma and Wang, 2013; Ma et al., 2020). High content of
Wuhan, 430071, China.
WCP, however, will significantly reduce the mechanical strength of
E-mail addresses: jsli@whrsm.ac.cn (J.-s. Li), zhaoyaqin@cumt.edu.cn (Y. Zhao).

https://doi.org/10.1016/j.envpol.2021.116776
0269-7491/© 2021 Elsevier Ltd. All rights reserved.
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

hydration products (Kim, 2017; Zhu et al., 2016). After high tem- Table 1
perature treatment, dehydrated WCP can be utilized to produce Proportions of components in different mixes of mortar and UCS.

dehydrated cement paste (DCP) and cement clinker (Florea et al., Motar Cement(kg) Sand (kg) Water (kg) 28d UCS (Mpa)
2014; Kwon et al., 2015). However, this recycling method con- M0.65 1 2.51 0.65 29.16
sumes high energy and causes great carbon emission. Besides, WCP M0.5 1 1.65 0.5 35.44
also be used as a silica source to synthesize geopolymer binder and M0.39 1 1.09 0.39 44.97
for the extraction of calcium-containing compounds to sequester M0.35 1 0.91 0.35 49.87

CO2 (Mun and Cho, 2013; Ren et al., 2020; Van der Zee and Zeman,
2016).Despite the availability of the above technologies for recy-
cling WCP, large-scale application is still limited due to its hetero- groups with different particle size ranges: 2-1 mm, 1e0.5 mm,
geneous properties and high hygroscopicity. 0.5e0.25 mm, 0.25e0.1 mm and <0.1 mm. Further, mortars were
Apart from reusing as construction materials, WCP has the po- heated in a muffle furnace at 200, 350, 600, and 850  C respectively

tential of being used as an adsorption material due to its alkalinity (rate: 10 C/min) for 60 min. The chemical composition of the
and complex pore structure, which promote the adsorption of cement, WCP and TMWCP is presented in Table 2.The procedure to
heavy metal ions in solution on the surface of the hydration com- obtain TMWCP is presented in Fig. 1.
ponents. Compared to other adsorbents such as modified bio-
char(Sharma and Naushad, 2020; Wan et al., 2019), ion
2.2. Batch adsorption experiment
exchangers(Naushad et al., 2015), and nanomaterials(Naushad,
2014), WCP is more economical. Some studies reported the
The stock Pb (II) solution (1000 mg/L) was prepared by dis-
adsorption of Cu(II), Zn(II), Pb (II) onto crushed concrete fines
solving analytical grade Pb (NO3)2 in deionized water, and the
(CCF)in solution by surface precipitation, showing that Pb (II)
desired concentration of Pb (II) solutions were obtained by diluting
could diffuse into the cement matrix and captured(Coleman et al.,
the stock solution. All adsorption experiments were performed by
2005). Furthermore, Kumara studied the adsorption of Cd (II) and
mixing a specific amount of TMWCP with 80 mL lead-containing
Pb (II) on autoclaved aerated concrete (AAC) fines and CCF that was
solution in a 100 mL polyethylene tube which rotated at 180 rpm
similar to WCP, and reported that AAC’s higher porosity made it a
in a thermostatic shaker (SHA-B300) to promote reaction.
better adsorbent than CCF. The main mechanisms of adsorption are
After reaction, the suspension was centrifuged and filtered with
ion exchange with the calcium in AAC, surface complexation and
a 0.45 mm aqueous filter. The metal concentrations were tested by
precipitation(Kumara et al., 2019). Dos Reis studied the effect of
inductively coupled plasma/optical emission spectrometry (ICP-
carbonation on WCP’s phosphorus adsorption capacity and found
OES PerkinElmer Optima 5300 DV). The amount of lead retained by
that carbonization inhibited the release of calcium and thus hin-
TMWCP was determined from equation (1).
dered phosphorus’s surface precipitation(dos Reis et al., 2020).
Moreover, other researchers practiced thermal modification to ðc0  ce ÞV
promote hydroxyapatite formation aiming at enhancing the qe ¼ (1)
m
adsorption of phosphorus. (Liu et al., 2020). Previous studies have
demonstrated satisfactory performance of WCP in adsorbing heavy where qe is the equilibrium adsorption quantity (mg/g), c0 and ce
metals, it is still not clear how the matrix and physicochemical are the initial and equilibrium lead concentration (mg/L), V is the
properties of WCP affect such performance. Besides, it is desirable solution volume (L), and m is the mass of WCP or TMWCP (g).
to understand the effect of thermal treatment on the adsorption
capacity of WCP. This study aimed to gain insight into the effec-
2.2.1. The selection of TMWCP
tiveness of thermally modified waste concrete powder (TMWCP),
The lead removal performance of different TMWCP samples
as an adsorbent in removal of Pb (II) from aqueous solutions. The
synthesized at different conditions was compared to identify the
specific objectives include a) to assess the effect of water-cement
best condition for producing TMWCP. Individual samples of 1.0g
ratio and particle size of cementitious products and subsequent
were added into 80 mL lead-containing solutions (1000 mg/L) with
thermal treatment on their performance in adsorption of Pb (II) via
an initial solution pH of 6.0 then allowed to react for 24 h at 25  C.
batch adsorption experiments; and b) to explore the adsorption
mechanisms through microscopic characterization techniques.
2.2.2. Influence of pH, ionic strength and dosage
2. Materials and methods The solution pH was adjusted by 0.1 mol/L HCl and NaOH and
measured by a pH meter (PHS-3E). The tests on pH influence were
2.1. Preparation of TMWCP carried out by adding 1.0g TMWCP to 80 mL lead-containing so-
lution of different initial pH values (1.00e6.00) at 25  C. Consid-
Considering the wide and heterogeneous sources of WCP, the ering the practical application and precipitation may be one of the
simulated cement mortars were produced in the laboratory to removal mechanisms of Pb (II), this study only controlled the initial
better study the removal characteristics of heavy metals, which was pH and recorded the final pH of solution. The same heavy metal
also adopted by other studies of WCP (Kwon et al., 2015; Volchek concentration (1000 mg/L) was used as the previous study to
et al., 2011). Briefly, ordinary Portland cement, river sand and wa- compare the adsorption characteristics(Coleman et al., 2005).
ter were mixed at different water-cement ratios of 0.65, 0.50, 0.39 Ca(NO3)2 and NaNO3 solutions were used for exploring the ef-
and 0.35 respectively (labeled as M0.65, M0.5, M0.39, M0.35), fect of ionic strength on Pb (II) adsorption. The tests were carried
which were in line with construction projects in China. The mixes out by adding 1.0 g TMWCP to 80 mL lead-containing solution
were then cast into square molds with a side length of 100 mm, and (1000 mg/L) containing Ca(NO3)2 and NaNO3 of different concen-
cured under 20  C, 98% humidity. After 28 days, the unconfined trations (0.05e2 mol/L) at 25  C.
compressive strength (UCS) of the samples were tested according The effect of dosage on adsorption was evaluated by altering the
to ASTM C109/109M (ASTM, 2013). The mix proportions and the solid-liquid ratio (2e50 g/L): different amounts of TMWCP were
UCS results of the four cement mortar mixes are shown in Table 1. added into 80 mL of lead-containing solution (1000 mg/L) with
The mortars were then crushed, sieved and categorized into five initial pH 6.0 at 25  C.
2
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Table 2
Chemical compositions of cement, WCP and TMWCP.

Sample Mass fraction (%)

SiO2 Al2O3 Fe2O3 CaO MgO SO3 K2O Na2O Others LOI

Cement 19.32 4.08 4.41 61.08 2.44 3.86 0.86 1.05 2.9 1.3
WCP 27.66 6.25 5.69 51.64 2.40 1.53 1.60 0.11 3.12 2.1
TMWCP 33.04 6.85 4.66 45.86 3.21 2.16 1.60 0.18 2.44 0.37

WCP: water-cement ratio 0.35.


TMWCP: water-cement ratio 0.35, modification temperature 850  C.

Fig. 1. Production procedures of TMWCP in this study.

2.2.3. Adsorption kinetics and isotherms spectroscopy (FTIR Nicolet Nexus 470) in the mid-infrared region
Adsorption kinetics was studied under initial Pb (II) concen- (400-4000 cm1) was performed to identify the spectral feature.
tration of 500 mg/L and 1000 mg/L with initial pH 6.0 at 25  C and XPS was conducted using a photoelectron spectrometer (ESCALAB
then taken out to measure the concentration of Pb (II) in the so- 250, Thermo Fisher Scientific) with monochromatic Al Ka radiation.
lution at intervals within 24 h. Data were fitted with the pseudo- The pass energy for narrow scans was 30 eV for fine spectrum and
first-order kinetic model, pseudo-second-order kinetic model and 100eV for full spectrum. The binding energy reference on the C 1s
intraparticle diffusion model for evaluation. line at 284.8 eV. Shirley-Linear background correction and
Isotherms were studied under the temperature of 25  C, 35  C, GaussianeLorentzian fitting were used for peak separation. The
and 45  C, respectively. Briefly, 1.0g TMWCP was added to 80 mL of pore size distribution and specific surface area of the TMWCP were
lead solutions of different concentrations (5e1500 mg/L) and determined by the BET test (Anton Paar NOVA1000e) with nitrogen
allowed to react for 24 h. Data were fitted into the Langmuir model adsorption. Samples are degassed in advance at 90  C for 12h under
and Freundlich for evaluation (shown in Appendix Text A.1). All the vacuum.
experiments were conducted in duplicate and the arithmetic mean
values were adopted. Additional tests were carried out when the 3. Results and discussion
difference lager than 5.0%.
3.1. Adsorption effect
2.3. Microstructural and spectroscopic analysis
3.1.1. Effect of WCP type
The TMWCP samples were analyzed by X-ray diffraction (XRD, As real WCP vary in particle sizes and water-cement ratio in the
AXS D8 Advance) with an interval of 5 /min at a 2q angle in the original mixes, the influence of these two parameters on the
range of 10 e80 to identify the mineral composition and chemical adsorption performance of the TMWCP was studied. As shown in
composition. Scanning electron microscope and energy dispersive Fig. 2(a), the adsorption capacity of Pb (II) by WCP increases with
X-ray spectrometer (SEM-EDX FEI Quanta250) were used to the water-cement ratio. The better adsorption capacity is probably
examine the surface morphology and element distribution in related to the more complex pore structure of WCP with high
TMWCP after lead adsorption. Fourier transform infrared water-cement ratio: the presence of excess water in cement paste
3
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Fig. 2. (a) Effect of water-cement ratio and particle size on WCP (b) Effect of temperature and water-cement ratio of WCP and TMWCP (0.5e0.25 mm).

distanced the cement grains and caused rapid chemical reactions 0.35 heated up to 850  C and fell in size range of 0.5e0.25 mm was
followed by growth of large crystals for hydration products, which adopted for investigating factors influencing Pb (II) adsorption.
subsequently contributed to an increase in pore size of the
matrix(Aïtcin, 1998; Piasta and Zarzycki, 2017). Furthermore, 3.1.2. Effect of initial pH
TMWCP from thermally WCP with high water-cement ratio has The solution pH has a significant impact on the uptake of heavy
more pores due to the dehydration, which explains the superior metals by the adsorption material, since it affects the surface
adsorption capacity of TMWCP with high water-cement ratio. The charges of the adsorbent and the heavy metals ions. No experiment
study by Pakravan et al. (2014) proved that mortars with high at solution pH > 6.0 was conducted to avoid precipitation of Pb (II)
water-cement ratio had higher surface free energy, which un- (Kazak and Tor, 2020). The simulation results of MINTEQ software
doubtedly increased their adsorption capacity as adsorption occurs (version 3.1) confirmed that Pb (II) is the main species of lead in
at the interface. Moreover, water enhances the shrinkage of the system (90.9%) at pH 6.0 in this study. Fig. 3 shows the effect of pH
mortar, resulting in cracking then reinforced Pb (II) removal. From on the adsorption performance of TMWCP. The adsorption capacity
the perspective of temperature effect as shown in Fig. 2(b) the (Qe) increases from 21.9 mg/g (pH ¼ 1.0) to more than 60 mg/g
adsorption capacity of TMWCP was found higher compared with (pH ¼ 2e6), whereas the removal rate increases from 29.3%
WCP regardless of water-cement ratio. After high temperature (pH ¼ 1.0) to more than 80% (pH ¼ 2.0e6.0).Meanwhile, the final
modification, the adsorption capacity of WCP produced with water- pH of the solution changes from acidic (pH ¼ 1.5) to alkaline
cement ratio of 0.35 increased by 36.8%, while the increase was (pH ¼ 11.7). The results can be explained by considering the elec-
only 7% in WCP produced with water-cement ratio of 0.65. It can be trostatic interaction between the surface charge of TMWCP and
concluded that the WCP produced with low water-cement ratio is lead ions. WCP after thermal modification and rehydration reaction
more sensitive to thermal modification, which enhances the contains various metal oxides which make its aqueous solution
applicability of TMWCP, namely, the low water-cement ratio WCP alkaline(Ma and Wang, 2013). Portlandite produced by the WCP
can achieve a lead removal capability similar to the high-water-
cement ratio WCP through thermal modification. The effect of
thermal modification on adsorption will be discussed in the section
of adsorption mechanism.
On the effect of particle size, TMWCP samples with particle size
in the range of 0.25e0.5 mm exhibited higher adsorption capacity
irrespective of the water-cement ratio. Generally speaking, due to
difference in hardness of cement and aggregate in the mortar,
smaller TMWCP particles collected after crushing have higher
specific surface areas and contain more cement paste (Pakravan
et al., 2014; Schoon et al., 2015). Nevertheless, this also means
more considerable surface energy, causing cohesion of TMWCP to
reduce the surface energy, the pores of the particles are filled with
rehydration products, which cripples the adsorption capacity. On
the contrary, it is generally accepted that the reduction in specific
surface area due to large particle size leads to a decrease in
adsorption capacity. It was found that TMWCP exhibited a satis-
factory lead removal function over a wide range of particle size and
modification temperature. Table S1 presented this section’s data to
make the effect of the three factors on Pb (II) removal clearer.
Considering the wider adsorption of low water cement ration in
practice, standardized TMWCP produced with water-cement ratio Fig. 3. Effect of initial pH on Pb (II) adsorption by TMWCP and the final pH of the
solution.

4
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

hydration decomposes into calcium oxide at high temperatures


then carbonizes into calcite, which continues to decompose then
produces calcium hydroxide after rehydration. A typical hydration
of WCP and rehydration of TMWCP proceed with the following
equations:a

C3 S þ nH2 O ¼ XCaO , SiO2 , yH2 O þ ð3  XÞCaðOHÞ2 (2)

CaðOHÞ2 / CaO þ H2 O (3)

CaO þ CO2 / CaCO3 (4)

CaCO3 / CaO þ CO2 (5)

CaO þ H2 O / CaðOHÞ2 (6)


The protons consume the alkaline functional groups of TMWCP
resulting in the inhibition of surface precipitation of lead ions in the
competition. The dramatic increase in the final pH when the initial
pH is increased from 1.0 to 3.0 represents the consumption of the Fig. 4. Effect of initial dosage of TMWCP on Pb (II) adsorption.
alkaline components in TMWCP in the acidic environment. The
hydration products in TMWCP present a surface charge due to the
increased from 86.03 mg/g to 92.66 mg/g. A further increase in
formation of silanol sites carried by the bridging silicate tetrahedra
dosage to 50 g/L resulted in a decrease in Qe. Pb (II) removal rate
or by the end-chain tetrahedra when the chains shorten(Pointeau
greatly increased from 17.43% to 92.96% with an increase in
et al., 2006; Viallis-Terrisse et al., 2001). Equations (7) and (8)
adsorbent dosage from 2 to 10 g/L. Thereafter, the removal rate
explain the principle of surface charging:
remained steady and then slightly decreased. It was found that the
SiOH % SiO þ Hþ (7) highest lead ion removal rate was attained with dosage of 10 g/L in
this study. This finding is different from those reported in most
studies (Niu et al., 2020; Wang et al., 2019). It can be explained that
SiOH þ Ca2þ % SiOCaþ þ Hþ (8) greater amount of TMWCP provide more active sites for Pb (II)
The calcium ions dissolved in the solution of TMWCP have a adsorption in the range from 2 g/L to 10 g/L. Beyond the upper
robust physical affinity with the surface of silicate, which causes threshold, the adsorption amount decreased sharply and the
excessive calcium adsorption on the surface of TMWCP making it removal rate drop slightly. This phenomenon was related to the
positively charged. Accompanied by the increase in pH, the hy- increase in calcium ion in the solution as a result of calcium
droxide ion neutralizes the surface charge of TMWCP, which in turn dissolution in the unhydrated component (Wang et al., 2015). The
enhances Pb (II) attraction. Obviously, the removal of Pb (II) by study on the effect of pH indicate that a high concentration of
TMWCP could be achieved under the condition that the final pH of calcium ion inhibits TMWCP from adsorption by altering the ma-
the solution was acidic, indicating some removal mechanisms terial’s surface potential. Hence, high dosage increased the active
except precipitation. pH results indicate that there is a strong sites in the system, but brought more calcium ions and impaired
bonding reaction between Pb (II) and TMWCP that can overcome the adsorption performance. Furthermore, boundary nucleation
the electrostatic repulsion since this bonding reaction probably occurred on the surface of excessive TMWCP in the solution, which
result in coordinate-covalent bonding forming between Pb (II) and held the rehydration products causing aggregation of TMWCP. Such
TMWCP surface functional groups which are inner-sphere complex aggregation would reduce the total porosity and hinder the diffu-
or surface precipitation. The Pb (II) adsorption capacities of several sion of Pb (II) into the adsorbent (Wen et al., 2018).
solid waste-based adsorbents at similar pH conditions are dis-
played in Table 3. Compared with other adsorbents, the TMWCP has 3.1.4. Effect of ionic strength
remarkable adsorption capacity in acid solution and more oppor- Changing the ionic strength of the solution can alter the
tunity for industrial applications, because most industrial waste- adsorption process in at least two ways: (a) affecting the adsor-
water containing heavy metals is acidic. bent’s surface potential and changing the current form of ions. (b)
Competing with the heavy metal ions and transforming the elec-
3.1.3. Effect of dosage trostatic interactions among the ions(Vilar et al., 2005). The effect
The influence of dosage on Pb (II) removal is illustrated in of ionic strength on the adsorption of Pb (II) in TMWCP is presented
Fig. 4.With an increase of dosage from 2 to 10 g/L, the Qe slightly in Fig. 5. At the same initial Pb (II) concentration, the adsorption

Table 3
Comparison of observed Qm for Pb (II) with different types of solid waste adsorbents.

Adsorbent S:L Ratio pH Qm(mg/g) Reference

TMWCP 1:80 6.00 66.59 This study


Fraction of crushed concrete 1:40 Not control 37.00 Coleman et al. (2005)
Waterworks sludge 3:500 5 20.40 Faisal et al. (2020)
Magnetic hydrochar 1:2000 5.5 183.37 Kazak and Tor (2020)
Incinerated sewage sludge ash 3:400 4 58.28 Wang et al. (2019)
Functional biochar 1:4000 5 133.60 Yin et al. (2019)

5
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Fig. 5. Effect of ionic strength on Pb (II) adsorption by TMWCP.


Fig. 7. Adsorption isotherm of Pb (II) by TMWCP with C0 in the range of 0e1500 mg/L.

quantity decreased sharply from 66.7 mg/g to about 33.8 mg/g


when the concentration of Ca(NO3)2 was increased from 0.05 to probably caused by their hindrance to the ion exchange between
2 mol/L and the Pb (II) removal rate dropped from 93.0% to around calcium in the TMWCP and lead ions in solution. On the other hand,
50.7%. The effect was very different with the addition of NaNO3. The the mild effect of sodium ion on Pb (II) adsorption suggests that
adsorption quantity increased slightly when the concentration of inner-sphere, rather than outer-sphere complexation is the domi-
NaNO3 increased from 0.05 to 0.5 mol/L then remained steady. nant means of adsorption (Lützenkirchen, 1997), because the
Studies have revealed that increasing the concentration of ions in chemical bonds formed in inner-sphere complexation are generally
the solution would reduce the thickness of the double layer and more stable and less susceptible to electrostatic interaction be-
shield the high-charged surface complexes of the TMWCP causing tween ions. The slight beneficial effect from NaNO3 may be ascribed
adverse effect on the adsorption (Wei et al., 2021). The inhibitory to the phenomenon of salting-out: the addition of an electrolyte
effect of calcium ions on the adsorption of Pb (II) in the TMWCP was affecting the solubility of a solute(Mehringer et al., 2020). Simula-
tion results from MINTEQ software (version 3.1) confirmed the

Fig. 6. Adsorption kinetics of Pb (II) by TMWCP:(a) pseudo-first-order model and pseudo-second-order model (b) intraparticle diffusion model.

Table 4
Results determined from adsorption kinetic models of Pb (II) adsorption by TMWCP.

C0(mg/L) qe,exp(mg/g) pseudo-first-order pseudo-second-order

qe,cal(mg/g) k1 R2 qe,cal(mg/g) k2 R2

500 47.63 45.01 2.06  102 0.98 49.15 5.51  104 0.99
1000 73.83 70.69 1.69  102 0.97 77.60 2.80  104 0.99

qe,exp: the experimental adsorption capacity.


qe,cal: the calculated adsorption capacity.

6
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Fig. 8. Adsorption models fitting results of Pb (II) adsorption by TMWCP at 25, 35 and 45  C with C0 ¼ 0e500 mg/L:(a) Langmuir model (b) Freundlich model.

Table 5
Results determined from Langmuir and Freundlich models of Pb (II) adsorption by TMWCP with C0 ¼ 0e500 mg/L.

Temperature Langmuir(0 < C0 < 500 mg/L) Freundlich(0 < C0 < 500 mg/L)

qm b R2 Kf n R2

25  C 46.02 0.46 0.99 1.237 1.33 0.93


35  C 38.58 0.02 0.99 0.762 1.18 0.96
45  C 30.01 0.03 0.98 0.763 1.38 0.94

existence of Pb(OH)3-(35.0%) and Pb3(OH)2þ


4 (37.5%) in the solution. adsorption controlling mechanism(Hamadi et al., 2001). Fig. 6(a)
The reactions are represented by Equations (9) and (10): presents the adsorption kinetic curves of TMWCP at initial con-
centrations of 500 mg/L and 1000 mg/L, the adsorption of Pb (II)
Pb2þ þ 3H2 O%PbðOHÞ3  þ 3Hþ (9) was rapid within the first 60min and then slowly attained equi-
librium at about 1440min. The adsorption quantity at 1000 mg/L
was greater than that at 500 mg/L, proving that a high concentra-
3Pb2þ þ 4H2 O % Pb3 ðOHÞ4 2þ þ 4Hþ (10)
tion of lead ions promoted adsorption, probably due to increased
Hence, sodium ions, with higher solubility, competes with lead probability of contact between lead ions and active surface sites.
ions for solvent molecules and inhibit the hydration of lead ions Lagergren’s pseudo-first-order adsorption kinetic model, Ho’s
thus strengthening the combination of adsorbate. Nevertheless, the pseudo-second-order adsorption kinetic model and intraparticle
salting-out effect has a limit beyond which the effect on Pb (II) diffusion model were used to check the fitness of the data using the
adsorption becomes steady. least square method (Ho et al., 1996; Lagergren, 1898; Pandey et al.,
2015). Table 4 gives the fitting results. The pseudo-second-order
3.1.5. Adsorption kinetics kinetic model was found more appropriate in predicting the
The investigation of adsorption kinetics is conducive to under- experimental data, indicating that the dominant controlling factor
standing the speed of adsorption and identify the primary of Pb (II) adsorption is the chemical effect associated with ion
exchange(Ho et al., 2011). The lower k2 value indicated that lead
was relatively stable on the adsorbent under this condition. There
are three main stages of adsorption: (1) lead ions moved to the
liquid film around TMWCP due to different concentration (2) lead
ions diffuse in the internal pores (3) lead ions combine with active
sites. The kinetic data were divided into three parts and checked by
the internal diffusion model separately (Fig. 6 (b)). The region with
a steeper slope at 1000 mg/L shows that a higher concentration
enhances the mass transfer rate in the system. The better fitting in
the 15e120min region with different concentrations reflects that
the determinant of the reaction rate became the internal diffusion
of the adsorbate. Finally, the gentle rise in the subsequent long
period shows that the combination of Pb (II) with active sites is the
determinant of the adsorption speed.

Fig. 9. Adsorption of Pb (II) on the surface of TMWCP(a) Monolayer adsorption at low 3.1.6. Adsorption isotherm
pH and concentration(b) Multi-layer adsorption at high pH and concentration. The shape of adsorption isotherms generally reflects the
7
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Fig. 10. SEM images and EDX analysis of WCP (a)(c), 850  C TMWCP (b)(d),850  C TMWCP after adsorption (e)(f).

8
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

characteristic of adsorbent surface and the interaction with adsor-


bates, which plays an essential role in the exploration of adsorption
mechanism. The isotherm of Pb (II) adsorption in the TMWCP at
different temperatures are presented in Fig. 7. The shape of the
tendency shows an atypical trend. As the initial concentration
increased, the equilibrium concentration of solution first raised
then swiftly declined, proving that the sudden increase of TMWCP’s
adsorption capacity at a certain concentration brought about a
higher Pb (II) adsorption amount. Similar phenomena have been
observed in the adsorption of wood extracts in bentonite at higher
initial concentrations and also lead adsorption in bentonite modi-
fied by sodium polyacrylate(Chen et al., 2020b; Heier et al., 2015).
The adsorption isotherms in low concentration zones (C0 < 500 mg/
L), however, show a typical trend. The linear form of Langmuir and
Freundlich models were used for checking data fitting with results
shown in Fig. 8, and related parameters listed in Table 5. Based on
the correlation coefficient (R2), the experimental data fit in the
Langmuir model better indicating that the adsorption of Pb (II) in
Fig. 11. Differential thermal analysis curve of M 0.35(30  Ce1100  C). TMWCP is mainly single-layer adsorption, which is consistent with
the findings of Kumara et al. (2019). The saturated adsorption ca-
pacities of TMWCP determined from the Langmuir adsorption
model at 25, 35, and 45  C are 46.02 mg/g, 38.58 mg/g and
30.01 mg/g respectively. The equilibrium capacity decreased with
increase in temperature, which confirms that the adsorption re-
action is exothermic. Moreover, the Langmuir model assumes only
one type of adsorption site on the surface with uniform energy
distribution(Liu, 2015). However, this simple assumption is not
necessarily valid in actual situation. In liquid phase adsorption
systems, the interaction between solute and solvent and the
interaction between adsorbent and solvent cause deviation to the
Langmuir model in some concentration ranges. Generally speaking,
the kick point of the adsorption isotherm indicates a change in
adsorption type(Zhang et al., 2019). When the initial concentration
in the system increases, the competition between the lead ions and
solvent molecules for adsorption sites becomes greater. The hy-
dration of TMWCP continues to generate alkaline compounds,
causing the solution’s pH to rise, and more lead ions tend to form
hydrated products. The lead hydrate cations carry a higher positive
charge than Pb (II) and tend to form outer-sphere complex on the
surface of TMWCP in the form of aqua compounds. The adsorption
type then changes from single layer to multiple layers, enhancing
the adsorption capacity rapidly (Fig. 9). To better illustrate the ex-
Fig. 12. XRD patterns of WCP and TMWCP before and after Pb (II) adsorption.
istence of lead in the system, the MINTEQ software (version 3.1)
was used to calculate the principal existence of lead in the
adsorption system (C0 ¼ 500 mg/L, pH 4e11) and the results are
given in Table S2.

3.2. Characterization of the TMWCP before and after adsorption

3.2.1. SEM-EDX analysis


Fig. 10 displays the SEM images of WCP and TMWCP (850  C)
before and after Pb (II) adsorption. Hydration products were
observed in layers on the surface of WCP particles, and no cracks
were found. Higher magnification image shows that the surface of
WCP is covered by flocculent CeSeH gels with sizes about 10 mm,
which indicates full hydration reaction of cement (Fig. 10(c)).
TMWCP (850  C) image shows visible cracks on the smoother and
denser surface as a result of vaporization of water and emission of
CO2 in the decomposition of CeSeH, Ca (OH)2 and CaCO3 in WCP at
high temperatures (Fig. 11(b,d)). Moreover, the discrepant expan-
sion of sand and other components is also a main cause for the
formation of pores (Arioz, 2007). After adsorption of Pb (II), a large
amount of needle-like ettringite is observed on the surface of
TMWCP due to the rehydration reaction. Spherules were also found
Fig. 13. FTIR spectra of WCP and TMWCP before and after adsorption. adhering to the TMWCP surface after adsorption, which is
9
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Fig. 14. XPS spectral results of TMWCP (850  C) before and after adsorption (a) Full spectrum results (b) Fine spectrum results of Pb4f orbit.

Fig. 15. N2 adsorption-desorption isotherms (a) and BJH pore size distributions (b) of WCP and TMWCP heated up to different temperatures.

Table 6
Specific surface area and porosity of TMWCP obtained by BET test.

Sample Specific surface area m2/g Pore volume ml/g Average pore size nm micropore% mesoporous% macropore%

WCP 3.120 0.020 12.401 0.280 45.413 54.307


200  C 3.329 0.024 14.997 0.306 37.279 62.415
350  C 3.028 0.025 16.023 0.120 35.842 64.038
600  C 3.303 0.028 17.039 0.385 27.051 72.564
850  C 2.912 0.029 20.123 0.064 19.489 80.447
850oCAfter adsorption 5.184 0.009 14.230 1.860 86.95 11.19

confirmed to be made up of Ca, Pb, Si, and O elements by EDX re- affect the WCP pore structure undoubtedly. The different expansive
sults, suggesting Pb (II) might fuse into silicate in TMWCP. Flaky Ca degree of aggregates and cement, synergism with the recrystalli-
(OH)2 was not detected, suggesting that the Ca (OH)2 was zation of amorphous CeSeH with long silicate chains also lead to
consumed by Pb (II). cracking within the WCP. XRD tests were carried out to identify the
phase after adsorption and the result is displayed in Fig. 12. Por-
tlandite (Ca(OH)2 PDF72-0156, 2q ¼ 18.11,34.17) was not detected in
3.2.2. TG, XRD and FTIR analysis TMWCP heated to high temperature (600 and 850  C), indicating
Three stages of rapid mass reduction around 91.89  C, 447.47  C, that Ca(OH)2 had been decomposed (Vysvaril et al., 2014). Corre-
and 717.51  C of TG/TGA test as displayed in Fig. 11 corresponding to spondingly, Ca(OH) 2 was also not observed in the SEM images of
the transformation and decomposition of CeSeH and ettringite, the TMWCP samples after Pb (II) adsorption. When heating to
the dehydroxylation of portlandite, and calcite decomposition, 850  C, the characteristic peak of Calcite (CaCO3 PDF86-2341,
respectively(Alonso and Fernandez, 2004; Chen et al., 2020a). The 2q ¼ 29.18) disappeared, reflecting the mass loss at 717  C in TG
processes of dehydroxylation and decarburization during heating
10
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Fig. 16. Images of TMWCP (water-cement ratio 0.35, temperature 850  C) particle before and after Pb (II) adsorption under 350 times magnification: (a) before adsorption (b) after
adsorption.

test. This suggests that CaCO3 may not be involved in the removal of 3.2.3. XPS analysis
Pb (II) as TMWCP without CaCO3 can adsorbs more Pb (II). The To verify the interaction between Pb (II) and TMWCP, the XPS
amorphous CeSeH with disordered layered structure and long spectrum analysis was conducted and the results are presented in
silicate chain could not be identified but was recrystallized into Fig. 14. After adsorption, the full spectrum of TMWCP emerged
dicalcium silicate(b-Ca2SiO4 PDF86-0398, 2q ¼ 32.94,39.45, peaks near 138.45 and 143.35 eV for Pb 4f7/2 and Pb 4f5/2,
41.93)under 850  C(Alonso and Fernandez, 2004). After Pb (II) respectively, suggesting the Pb (II) adsorption by TMWCP. Simul-
adsorption, characteristic peaks of Alamosite (PbSiO3 PDF29-0782, taneously, the intensity of the Ca 2p3/2 and Ca 2p1/2 peaks at
2q ¼ 13.72,43.92) appeared in the TMWCP, indicating that ion ex- binding energies of 347.5 eV and 351.5 eV decreased significantly
change occurred between Pb (II) and calcium silicate. after adsorption, inferring ion exchange reaction in the Pb (II)
FTIR spectra of the selected samples shown in Fig. 13 are in removal. The fine spectrum of Pb 4f is presented in Fig. 14(b) to
agreement with the XRD results. The peak at 3641 cm1 originating better determine the chemical state of Pb on the TMWCP surface.
from the eOH bond in Ca(OH)2 disappears after heating to 600  C Binding energies of 138.8 eV and 138.4 eV were found to be
and then appears after further heating to 850  C probably due to attributed to PbSiO3 and Pb(OH)2, respectively(Wang et al., 2021),
the rehydration of TMWCP (Li and Zhang, 2019). Similar to the XRD which confirmed two different mechanisms of the interaction be-
results, no OeH bond was detected in the TMWCP after lead tween Pb and TMWCP. Combined with the reduction of calcium
adsorption. Ca(OH)2 in WCP decomposes under 600  C and reacts after adsorption in the full spectrum, the formed PbSiO4 was
with CO2 then decomposed at 850  C, which is reflected in the FTIR attributed to ion exchange, as shown in equation (11). The precip-
results as change in intensity of OeCeO tensile vibration at 1425- itation was caused by the alkaline environment of pore solution, as
1437 cm1(Moraes et al., 2015; Yang et al., 2014). TMWCP (850  C) given in equation (12). Therefore, the XPS results are in high
exhibited a typical signal of asymmetric stretching of SieO at concordance with the XRD and FTIR results.
960 cm1(García Lodeiro et al., 2009; Li et al., 2020), which corre-
sponds to the presence of b-dicalcium silicate also detected by XRD. CaSiO3 þ Pb2þ / PbSiO3 þ Ca2þ (11)

CaðOHÞ2 þ Pb2þ / PbðOHÞ2 þ Ca2þ (12)

3.2.4. BET analysis


To evaluate the effect of thermal modification on the pore
structure of WCP, BET tests were conducted using nitrogen gas and
results of adsorption isotherms are presented in Fig. 15(a). The N2
adsorption-desorption isotherm curves of TMWCP conform to type
III of International Union of Pure and Applied Chemistry (IUPAC),
which means that N2 molecules were adsorbed in multiple layers
on the surface of TMWCP, that is, the adsorption energy of the nth
layer of TMWCP is much higher than the adsorption energy of the
first layer (En > E1). With increase in modification temperature, the
N2 adsorption capacity of TMWCP in the low-pressure zone de-
creases, but the final adsorption capacity increases. Conversely, the
adsorption capacity of the TMWCP after adsorption is consistently
low, both in the low and high-pressure regions, which is due to the
plugging of pores by the adsorbed products. The specific surface
areas and pore distribution calculated from the BJH formula based
on the N2 adsorption data are shown in Table 6 and Fig. 15(b)
respectively. The total specific surface area of TMWCP fluctuates
Fig. 17. Increase in Ca(II) concentration after Pb (II) adsorption by TMWCP modified with the temperature of modification as a result of the structural
with different temperatures. change of CeSeH, the decomposition and recrystallization of
11
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

hydration products(Zhang et al., 2013). This indicates that the in- buffering capacity for capturing Pb (II) to form a precipitate
crease in Pb (II) adsorption capacity of TMWCP is not related to the (Holmes et al., 2017). Pb (II) precipitation explains the absence of
increase in the specific surface area. High modification temperature Portlandite in XRD pattern and the disappearance of OeH peaks in
increases the number of macropores (>50 nm) and provides a FTIR tests, which should exist after TMWCP rehydration (Carriço
larger effective adsorption specific surface area (Georgali et al., et al., 2020).
2005), thereby enhancing Pb (II) adsorption capacity. The adsor-
bed products partially blocked the macropores, causing a conver- 3.3.2. Ion-exchange
sion from macropores to mesopores, which explains the sudden Ion-exchange interactions have also been found between cal-
increase in mesopores proportion of TMWCP after adsorption. The cium and lead. Recent studies have proved that the capture of heavy
digital microscope images also confirm the effectiveness of mac- metals by cementitious materials is strongly associated with the
ropores in adsorption (Fig. 16) as great amounts of adsorption replacement of calcium on the surface (Coleman et al., 2005). Pb (II)
product (yellow substance) are observed in the TMWCP cracks. can enter the CeSeH gel to replace some calcium to form
CePbeSeH as the ionic radius of lead and calcium are similar
3.3. Mechanisms of Pb (II) removal by TMWCP (1.17 Å and 1.00 Å, respectively) (Labhasetwar and Shrivastava,
1989). The formation of PbSiO3 as detected in the XRD tests and
The study of various factors affecting the adsorption of Pb (II) in the element composition identified in the EDX tests on the
the TMWCP has revealed the multiplicity of adsorption adsorption products proved the occurrence of ion exchange. In
mechanisms. addition to CeSeH, the chemical retention of Pb (II) by partial ex-
change of calcium and aluminum in the lattice can also be achieved
3.3.1. Precipitation by the ettringite observed in Fig. 10 (e)(f)(Contessi et al.,
The large amount of alkaline metal oxides produced by the 2021).Other studies reported similar mechanism in the adsorp-
TMWCP rehydration raised the solution pH to 12 (Fig. 3)(Ma and tion of Pb (II) in bentonite and other minerals (Chen et al., 2020c;
Wang, 2013), which promoted the precipitation of Pb (II) on the Meneguin et al., 2017).
surface of the TMWCP. BET results confirmed that high temperature
modification increased macropores which provided more space for 3.3.3. Complexation
precipitation. Besides, water within macropores after TMWCP The vibration at 1187 cm1 and 746 cm1 in the fingerprint area
rehydration provided essential source of alkalinity and self- of TMWCP after adsorption confirmed the formation of SieOeH

Fig. 18. Relationship between adsorption capacity of TMWCP and Ca(II) concentration (figures at upper right corners are water-cement ratio).

12
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

and SieOePb (Lazarev, 1972; Yin et al., 2019), while the silanol should be made to regeneration of TMWCP, pelleting of powdery
group was found to be significantly weakened in the TMWCP. This adsorbent for easy separation from water, as well as exploring the
phenomenon indicates the complexation process of adsorption, adsorption capacity for other pollutants. This finding sheds light on
which proceeds in two ways: the potential of large-scale application of TMWCP in industrial
(1) Lead ions replace the protons released by ionization of the wastewater treatment.
silanol group to complete the inner-sphere complexation (Equation
(13)). (2) The hydrated lead ions bond with the oxygen in the silanol Author statement
group by sharing electrons to complete the outer-sphere
complexation (Equation (14)). Jiang-shan Li, Conceptualization, Methodology, Writing e re-
view & editing. Zihan Ma, Methodology, Data curation, Writing e
2Si  O þ Pb2þ / Si  O  Pb  O  Si (13) original draft, Visualization. Runze Xue, Data curation, Visualiza-
tion. Yaqin Zhao, Supervision, Funding acquisition. Zhen Chen,
 2þ Writing e review & editing. Qiming Wang, Software. Chi Sun Poon,
Si  O þ PbðH2 OÞX / Si  O  H  OH  Pb2þ ðH2 OÞX
Conceptualization, Project administration.
(14)
Equations (7) and (8) illustrate the mechanism of surface charge Declaration of competing interest
on TMWCP. The presence of Ca (II) increases the positive charge on
the surface and competes with Pb (II) for complexation. High- The authors declare that they have no known competing
temperature modification restricts the dissolution of calcium on financial interests or personal relationships that could have
the surface of TMWCP probably due to structural changes in CeSeH appeared to influence the work reported in this paper.
and the formation of silicates with lower solubility and fluidity
(such as nesosilicate), thereby reducing its inhibition effect on Acknowledgement
adsorption. To demonstrate the inhibitory effect of Ca (II) on
complexation, the Ca (II) concentration in the aqueous solution This study is financially supported by the National Key Research
after Pb (II) adsorption in TMWCP synthesized from different pro- and Development Program (Grant No. 2019YFC1804002), National
cesses were tested, and the results are presented in Fig. 17. The Science Fund for Distinguished Young Scholars (Grant No.
release of calcium ions from TMWCP reduced with rise in modifi- 51625903), National Natural Science Foundation of China/Hong
cation temperature regardless of the water-cement ratio. It can be Kong Research Grants Council, University Grants Committee Joint
observed from Fig. 18 that the concentration of Ca (II) is highly Research Scheme (Grant No. 51861165104), Special Fund for Basic
correlated with the adsorption capacity of TMWCP. As such, Research on Science Instruments of the National Science Founda-
TMWCP from 850  C high temperature released the least Ca (II) and tion of China (Grant No. 51827814) and The CAS Pioneer Hundred
offered the highest Pb (II) adsorption capacity. Talents Program in China, Natural Science Foundation of Jiangsu
Province (Grant No. BK20160241), Postgraduate Research & Prac-
tice Innovation Program of Jiangsu Province (Grant No.
4. Conclusions SJCX20_0818), Assistance Program for Future Outstanding Talents
of China University of Mining and Technology (Grant No.
This study assessed deeper into the application of WCP on lead 2020WLJCRCZL055).
adsorption, rather than traditional landfilling or reuse in building
materials. WCP made from different water-cement ratios were Appendix A. Supplementary data
separated into groups of different particle sizes, and then heated
under different temperatures for producing materials to adsorb Pb Supplementary data to this article can be found online at
(II). The high water-cement ratio promoted adsorption by changing https://doi.org/10.1016/j.envpol.2021.116776.
the pore structure of the WCP. Simultaneously, TMWCP with
smaller particle size inhibited adsorption due to particle aggrega- References
tion induced by high surface energy. The results showed that high
temperature treatment significantly enhanced the adsorption ca- Aïtcin, P.-C., 1998. High Performance Concrete. CRC press.
pacity of WCP especially for those made with low water-cement Alonso, C., Fernandez, L., 2004. Dehydration and rehydration processes of cement
paste exposed to high temperature environments. J. Mater. Sci. 39 (9),
ratio from two mechanisms: (1) Enlarging the macropores of the 3015e3024.
WCP to provide adequate space for the precipitation of Pb (II) in the Arioz, O., 2007. Effects of elevated temperatures on properties of concrete. Fire Saf. J.
alkaline pore fluid; (2) Restricting the migration ability of calcium 42 (8), 516e522. https://doi.org/10.1016/j.firesaf.2007.01.003.
ASTM, A., 2013. Standard test method for compressive strength of hydraulic cement
from the surface of the WCP to the solution, enhancing the elec- mortars (using 2-in. or [50-mm] cube specimens). Ann. Book ASTM Stand. Ann.
tronegativity of TMWCP and thus reinforcing the surface Book ASTM Stand. 4 (1), 1e9.
complexation. The best modification temperature and particle size Carriço, A., Real, S., Bogas, J.A., Costa Pereira, M.F., 2020. Mortars with thermo
activated recycled cement: fresh and mechanical characterisation. Construct.
for Pb (II) removal was 850  C and 0.5e0.25 mm respectively. Batch Build. Mater. 256 https://doi.org/10.1016/j.conbuildmat.2020.119502.
adsorption experiments indicated that the initial pH, ionic strength Chen, X., Li, J., Xue, Q., Huang, X., Liu, L., Poon, C.S., 2020a. Sludge biochar as a green
of the solution and the dosage of TMWCP all influenced the additive in cement-based composites: mechanical properties and hydration
kinetics. Construct. Build. Mater. 262 https://doi.org/10.1016/
adsorption of Pb (II). The pseudo-second-order kinetic model and j.conbuildmat.2020.120723.
Langmuir model at low concentrations (C0 < 500 mg/g) predicted Chen, Y.-g., Liao, R.-p., Yu, C., Yu, X., 2020b. Sorption of Pb(II) on sodium polyacrylate
the adsorption process well. The maximum equilibrium adsorption modified bentonite. Adv. Powder Technol. https://doi.org/10.1016/
capacities of TMWCP at 25, 35 and 45  C were 46.02 mg/g, j.apt.2020.06.011.
Chen, Y., Wang, S., Li, Y., Liu, Y., Chen, Y., Wu, Y., Zhang, J., Li, H., Peng, Z., Xu, R.,
38.58 mg/g and 30.01 mg/g, respectively, but the values varied due Zeng, Z., 2020c. Adsorption of Pb(II) by tourmaline-montmorillonite composite
to the atypical adsorption isotherm of TMWCP at high concentra- in aqueous phase. J. Colloid Interface Sci. 575, 367e376. https://doi.org/10.1016/
tions. Microstructural analyses further confirmed that the adsorp- j.jcis.2020.04.110.
Coleman, N.J., Lee, W.E., Slipper, I.J., 2005. Interactions of aqueous Cu2þ, Zn2þ and
tion mechanisms mainly included surface precipitation, ion Pb2þ ions with crushed concrete fines. J. Hazard Mater. 121 (1e3), 203e213.
exchange and complexation. In the future researches, endeavor https://doi.org/10.1016/j.jhazmat.2005.02.009.

13
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

Contessi, S., Dalconi, M.C., Pollastri, S., Calgaro, L., Meneghini, C., Ferrari, G., 2020. Salting-in and salting-out effects of short amphiphilic molecules: a bal-
Marcomini, A., Artioli, G., 2021. Cement-stabilized contaminated soil: under- ance between specific ion effects and hydrophobicity. Phys. Chem. Chem. Phys.
standing Pb retention with XANES and Raman spectroscopy. Sci. Total Environ. Meneguin, J.G., Moise s, M.P., Karchiyappan, T., Faria, S.H.B., Gimenes, M.L., de
752, 141826. https://doi.org/10.1016/j.scitotenv.2020.141826. Barros, M.A.S.D., Venkatachalam, S., 2017. Preparation and characterization of
dos Reis, G.S., Thue, P.S., Cazacliu, B.G., Lima, E.C., Sampaio, C.H., Quattrone, M., calcium treated bentonite clay and its application for the removal of lead and
Ovsyannikova, E., Kruse, A., Dotto, G.L., 2020. Effect of concrete carbonation on cadmium ions: adsorption and thermodynamic modeling. Process Saf. Environ.
phosphate removal through adsorption process and its potential application as Protect. 111, 244e252. https://doi.org/10.1016/j.psep.2017.07.005.
fertilizer. J. Clean. Prod. 256 https://doi.org/10.1016/j.jclepro.2020.120416. Moraes, J.C.B., Akasaki, J.L., Melges, J.L.P., Monzo , J., Borrachero, M.V., Soriano, L.,
Duan, H., Miller, T.R., Liu, G., Tam, V.W.Y., 2019. Construction debris becomes Paya , J., Tashima, M.M., 2015. Assessment of sugar cane straw ash (SCSA) as
growing concern of growing cities. Waste Manag. 83, 1e5. https://doi.org/ pozzolanic material in blended Portland cement: microstructural character-
10.1016/j.wasman.2018.10.044. ization of pastes and mechanical strength of mortars. Construct. Build. Mater.
EPA, U., 2015. Advancing Sustainable Materials Management: 2014 Fact Sheet. 94, 670e677. https://doi.org/10.1016/j.conbuildmat.2015.07.108.
Faisal, A.A.H., Al-Wakel, S.F.A., Assi, H.A., Naji, L.A., Naushad, M., 2020. Waterworks Mun, M., Cho, H., 2013. Mineral carbonation for carbon sequestration with indus-
sludge-filter sand permeable reactive barrier for removal of toxic lead ions from trial waste. Energy Procedia 37, 6999e7005. https://doi.org/10.1016/
contaminated groundwater. J. Water Process Eng. 33 https://doi.org/10.1016/ j.egypro.2013.06.633.
j.jwpe.2019.101112. Naushad, M., 2014. Surfactant assisted nano-composite cation exchanger: devel-
Florea, M.V.A., Ning, Z., Brouwers, H.J.H., 2014. Activation of liberated concrete fines opment, characterization and applications for the removal of toxic Pb2þ from
and their application in mortars. Construct. Build. Mater. 50, 1e12. https:// aqueous medium. Chem. Eng. J. 235, 100e108. https://doi.org/10.1016/
doi.org/10.1016/j.conbuildmat.2013.09.012. j.cej.2013.09.013.
García Lodeiro, I., Macphee, D.E., Palomo, A., Ferna ndez-Jime nez, A., 2009. Effect of Naushad, M., Alothman, Z.A., Awual, M.R., Alam, M.M., Eldesoky, G.E., 2015.
alkalis on fresh CeSeH gels. FTIR analysis. Cement Concr. Res. 39 (3), 147e153. Adsorption kinetics, isotherms, and thermodynamic studies for the adsorption
https://doi.org/10.1016/j.cemconres.2009.01.003. of Pb2þ and Hg2þ metal ions from aqueous medium using Ti(IV) iodovanadate
Georgali, B., Tsakiridis, P.E.J.C., Composites, C., 2005. Microstructure of fire-damaged cation exchanger. Ionics 21 (8), 2237e2245. https://doi.org/10.1007/s11581-
concrete. Case Study 27 (2), 255e259. 015-1401-7.
Hamadi, N.K., Chen, X.D., Farid, M.M., Lu, M.G.Q., 2001. Adsorption kinetics for the Niu, M., Li, G., Cao, L., Wang, X., Wang, W., 2020. Preparation of sulphate aluminate
removal of chromium(VI) from aqueous solution by adsorbents derived from cement amended bentonite and its use in heavy metal adsorption. J. Clean.
used tyres and sawdust. Chem. Eng. J. 84 (2), 95e105. https://doi.org/10.1016/ Prod. 256 https://doi.org/10.1016/j.jclepro.2020.120700.
S1385-8947(01)00194-2. Pakravan, H.R., Jamshidi, M., Latifi, M., 2014. Relationship between the surface free
Heier, D., Blackstock, T., Stack, K., Richardson, D., Lewis, T., 2015. Adsorption of wood energy of hardened cement paste and chemical phase composition. J. Ind. Eng.
extractives and model compounds onto bentonite. Colloid. Surface. Phys- Chem. 20 (4), 1737e1740. https://doi.org/10.1016/j.jiec.2013.08.025.
icochem. Eng. Aspect. 482, 213e221. https://doi.org/10.1016/ Pandey, P.K., Sharma, S.K., Sambi, S.S., 2015. Removal of lead(II) from waste water on
j.colsurfa.2015.05.018. zeolite-NaX. J. Environ. Chem. Eng. 3 (4), 2604e2610. https://doi.org/10.1016/
Ho, Y.-S., Wase, D.A.J., Forster, C.F., 1996. Removal of lead ions from aqueous solution j.jece.2015.09.008.
using sphagnum moss peat as adsorbent. WaterSA 22, 219e224. Piasta, W., Zarzycki, B., 2017. The effect of cement paste volume and w/c ratio on
Ho, Y.S., Ng, J.C.Y., McKay, G., 2011. Kinetics of pollutant sorption by biosorbents: shrinkage strain, water absorption and compressive strength of high perfor-
review. Separ. Purif. Methods 29 (2), 189e232. https://doi.org/10.1081/spm- mance concrete. Construct. Build. Mater. 140, 395e402. https://doi.org/10.1016/
100100009. j.conbuildmat.2017.02.033.
Holmes, R.R., Hart, M.L., Kevern, J.T., 2017. Heavy metal removal capacity of indi- Pointeau, I., Reiller, P., Mace, N., Landesman, C., Coreau, N., 2006. Measurement and
vidual components of permeable reactive concrete. J. Contam. Hydrol. 196, modeling of the surface potential evolution of hydrated cement pastes as a
52e61. https://doi.org/10.1016/j.jconhyd.2016.12.005. function of degradation. J. Colloid Interface Sci. 300 (1), 33e44. https://doi.org/
Huang, B., Wang, X., Kua, H., Geng, Y., Bleischwitz, R., Ren, J., 2018. Construction and 10.1016/j.jcis.2006.03.018.
demolition waste management in China through the 3R principle. Resour. Ren, P., Li, B., Yu, J.-G., Ling, T.-C., 2020. Utilization of recycled concrete fines and
Conserv. Recycl. 129, 36e44. https://doi.org/10.1016/j.resconrec.2017.09.029. powders to produce alkali-activated slag concrete blocks. J. Clean. Prod. 267
Kaliyavaradhan, S.K., Ling, T.-C., Mo, K.H., 2020. Valorization of waste powders from https://doi.org/10.1016/j.jclepro.2020.122115.
cement-concrete life cycle: a pathway to circular future. J. Clean. Prod. 268 Schoon, J., De Buysser, K., Van Driessche, I., De Belie, N., 2015. Fines extracted from
https://doi.org/10.1016/j.jclepro.2020.122358. recycled concrete as alternative raw material for Portland cement clinker pro-
Kazak, O., Tor, A., 2020. In situ preparation of magnetic hydrochar by co- duction. Cement Concr. Compos. 58, 70e80. https://doi.org/10.1016/
hydrothermal treatment of waste vinasse with red mud and its adsorption j.cemconcomp.2015.01.003.
property for Pb(II) in aqueous solution. J. Hazard Mater. 393, 122391. https:// Sharma, G., Naushad, M., 2020. Adsorptive removal of noxious cadmium ions from
doi.org/10.1016/j.jhazmat.2020.122391. aqueous medium using activated carbon/zirconium oxide composite: isotherm
Kim, Y.-J., 2017. Quality properties of self-consolidating concrete mixed with waste and kinetic modelling. J. Mol. Liq. 310 https://doi.org/10.1016/
concrete powder. Construct. Build. Mater. 135, 177e185. https://doi.org/10.1016/ j.molliq.2020.113025.
j.conbuildmat.2016.12.174. Van der Zee, S., Zeman, F., 2016. Production of carbon negative precipitated calcium
Kumara, G.M.P., Kawamoto, K., Saito, T., Hamamoto, S., Asamoto, S., 2019. Evaluation carbonate from waste concrete. Can. J. Chem. Eng. 94 (11), 2153e2159. https://
of autoclaved aerated concrete fines for removal of Cd(II) and Pb(II) from doi.org/10.1002/cjce.22619.
wastewater. J. Environ. Eng. 145 (11) https://doi.org/10.1061/(asce)ee.1943- Viallis-Terrisse, H., Nonat, A., Petit, J.-C., 2001. Zeta-potential study of calcium sili-
7870.0001597. cate hydrates interacting with alkaline cations. J. Colloid Interface Sci. 244 (1),
Kwon, E., Ahn, J., Cho, B., Park, D., 2015. A study on development of recycled cement 58e65. https://doi.org/10.1006/jcis.2001.7897.
made from waste cementitious powder. Construct. Build. Mater. 83, 174e180. Vilar, V.J.P., Botelho, C.M.S., Boaventura, R.A.R., 2005. Influence of pH, ionic strength
https://doi.org/10.1016/j.conbuildmat.2015.02.086. and temperature on lead biosorption by Gelidium and agar extraction algal
Labhasetwar, N., Shrivastava, O.P., 1989. Ca2þ & Pb2þ exchange reaction of calcium waste. Process Biochem. 40 (10), 3267e3275. https://doi.org/10.1016/
silicate hydrate: Ca5Si6O18H2 .4H2O. J. Mater. Sci. 24 (12), 4359e4362. j.procbio.2005.03.023.
Lagergren, S., 1898. About the theory of so-called adsorption of solution substances. Volchek, K., Miah, M.Y., Kuang, W., DeMaleki, Z., Tezel, F.H., 2011. Adsorption of
Handlinger 24. cesium on cement mortar from aqueous solutions. J. Hazard Mater. 194,
Lazarev, A.N., 1972. Vibrational Spectra and Structure of Silicates. 331e337. https://doi.org/10.1016/j.jhazmat.2011.07.111.
Li, G., Zhang, L.W., 2019. Microstructure and phase transformation of graphene- Vysvaril, M., Bayer, P., Chrom a, M., Rovnaníkov a, P., 2014. Physico-mechanical and
cement composites under high temperature. Compos. B Eng. 166, 86e94. microstructural properties of rehydrated blended cement pastes. Construct.
https://doi.org/10.1016/j.compositesb.2018.11.127. Build. Mater. 54, 413e420. https://doi.org/10.1016/j.conbuildmat.2013.12.021.
Li, W., Mao, Z., Xu, G., Chang, H., Hong, J., Zhao, H., Xu, J., Liu, Z., 2020. Study on the Wan, Z., Cho, D.W., Tsang, D.C.W., Li, M., Sun, T., Verpoort, F., 2019. Concurrent
early cement hydration process in the presence of cationic asphalt emulsion. adsorption and micro-electrolysis of Cr(VI) by nanoscale zerovalent iron/
Construct. Build. Mater. 261 https://doi.org/10.1016/j.conbuildmat.2020.120025. biochar/Ca-alginate composite. Environ. Pollut. 247, 410e420. https://doi.org/
Liu, D., Quan, X., Zhu, H., Huang, Q., Zhou, L., 2020. Evaluation of modified waste 10.1016/j.envpol.2019.01.047.
concrete powder used as a novel phosphorus remover. J. Clean. Prod. 257 Wang, Q., Li, J.S., Poon, C.S., 2019. Using incinerated sewage sludge ash as a high-
https://doi.org/10.1016/j.jclepro.2020.120646. performance adsorbent for lead removal from aqueous solutions: perfor-
Liu, S., 2015. A mathematical model for competitive adsorptions. Separ. Purif. mances and mechanisms. Chemosphere 226, 587e596. https://doi.org/10.1016/
Technol. 144, 80e89. https://doi.org/10.1016/j.seppur.2015.01.044. j.chemosphere.2019.03.193.
Lützenkirchen, J., 1997. Ionic strength effects on cation sorption to oxides: macro- Wang, T., Zhang, P., Wu, D., Sun, M., Deng, Y., Frost, R.L., 2015. Effective removal of
scopic observations and their significance in microscopic interpretation. zinc (II) from aqueous solutions by tricalcium aluminate (C(3)A). J. Colloid
J. Colloid Interface Sci. 195 (1), 149e155. https://doi.org/10.1006/jcis.1997.5160. Interface Sci. 443, 65e71. https://doi.org/10.1016/j.jcis.2014.11.046.
Ma, X., Wang, Z., 2013. Effect of ground waste concrete powder on cement prop- Wang, Z., Jiang, Y., Mo, X., Gu, X., Li, W., 2021. Speciation transformation of Pb during
erties. Adv. Mater. Sci. Eng. 2013, 1e5. https://doi.org/10.1155/2013/918294. palygorskite sorption-calcination process: implications for Pb sequestration.
Ma, Z., Liu, M., Duan, Z., Liang, C., Wu, H., 2020. Effects of active waste powder Appl. Geochem. 124 https://doi.org/10.1016/j.apgeochem.2020.104850.
obtained from C&D waste on the microproperties and water permeability of Wei, X., Pan, D., Xu, Z., Xian, D., Li, X., Tan, Z., Liu, C., Wu, W., 2021. Colloidal stability
concrete. J. Clean. Prod. 257 https://doi.org/10.1016/j.jclepro.2020.120518. and correlated migration of illite in the aquatic environment: the roles of pH,
Mehringer, J., Hofmann, E., Touraud, D., Koltzenburg, S., Kellermeier, M., Kunz, W., temperature, multiple cations and humic acid. Sci. Total Environ. 768, 144174.

14
Z. Ma, R. Xue, J.-s. Li et al. Environmental Pollution 277 (2021) 116776

https://doi.org/10.1016/j.scitotenv.2020.144174. biochar-based functional materials derived from biowaste for Pb(II) removal.
Wen, X., Du, C., Zeng, G., Huang, D., Zhang, J., Yin, L., Tan, S., Huang, L., Chen, H., Appl. Surf. Sci. 465, 297e302. https://doi.org/10.1016/j.apsusc.2018.09.010.
Yu, G., Hu, X., Lai, C., Xu, P., Wan, J., 2018. A novel biosorbent prepared by Zhang, L., Chen, L., Huang, G., Liu, F., 2019. Gibberellic acid surface complexation on
immobilized Bacillus licheniformis for lead removal from wastewater. Chemo- ferrihydrite at different pH values: outer-sphere complexes versus inner-sphere
sphere 200, 173e179. https://doi.org/10.1016/j.chemosphere.2018.02.078. complexes. Sci. Total Environ. 650 (Pt 1), 741e748. https://doi.org/10.1016/
Xiao, J., Ma, Z., Sui, T., Akbarnezhad, A., Duan, Z., 2018. Mechanical properties of j.scitotenv.2018.09.023.
concrete mixed with recycled powder produced from construction and de- Zhang, Q., Ye, G., Koenders, E., 2013. Investigation of the structure of heated Port-
molition waste. J. Clean. Prod. 188, 720e731. https://doi.org/10.1016/ land cement paste by using various techniques. Construct. Build. Mater. 38,
j.jclepro.2018.03.277. 1040e1050. https://doi.org/10.1016/j.conbuildmat.2012.09.071.
Yang, Y., Pignatello, J.J., Ma, J., Mitch, W.A., 2014. Comparison of halide impacts on Zhu, P., Mao, X., Qu, W., Li, Z., Ma, Z.J., 2016. Investigation of using recycled powder
the efficiency of contaminant degradation by sulfate and hydroxyl radical-based from waste of clay bricks and cement solids in reactive powder concrete.
advanced oxidation processes (AOPs). Environ. Sci. Technol. 48 (4), 2344. Construct. Build. Mater. 113, 246e254. https://doi.org/10.1016/
Yin, W., Dai, D., Hou, J., Wang, S., Wu, X., Wang, X., 2019. Hierarchical porous j.conbuildmat.2016.03.040.

15

You might also like