Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53

Contents lists available at ScienceDirect

Process Safety and Environmental Protection

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

Kinetics and thermodynamics of U(VI) ions from


aqueous solution using oxide nanopowder

Mohamed Ahmed Mahmoud a,b,∗


a Nuclear Material Authority, Cairo, Egypt
b Faculty of Engineering, Jazan University, Jazan, Saudi Arabia

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

Article history: Aluminum oxide nanopowder (AONP) was used for the preconcentration and recovery of
Received 21 May 2015 uranium ions from an aqueous solution. Adsorption process in batch system was carried
Received in revised form 25 out by varying pH, initial U(VI) concentration, adsorbent dose, adsorption time and tem-
December 2015 perature. The adsorption efficiency could reach 99.85% at pH 5.0, 150 mg dose and 303 K.
Accepted 12 February 2016 Desorption of uranium ions can be carried out using 1.5 M HNO3 . Equilibrium adsorption was
Available online 21 February 2016 attained within 40 min at 303 K and within 20 min at 333 K indicating that the rate of U(VI)
uptake was found to be faster with increasing temperature. Adsorption data indicates the
Keywords: process following Langmuir isotherm and pseudo-second-order kinetic model. The mean
Adsorption energy, enthalpy, and activation energy confirming that the adsorption of U(VI) onto AONP
Recovery is physical adsorption. Moreover, the thermodynamic parameters showed the endothermic
U(VI) and spontaneous nature of the adsorption process.
Al2 O3 nanopowder © 2016 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Kinetics
Thermodynamics
Isotherms
Design

1. Introduction widely used for removal of heavy metals from aqueous medium. A lot of
adsorbent materials are being investigated for the removal of U(VI) ions
Uranium is released into the environment through numerous sources, from aqueous solutions including activated carbon (Abbasi and Streat,
such as nuclear industries, natural deposits, fertilizers and other 1994), dunite (Konstantin and Demetriou, 2007), akaganeite (Yusan and
processing of uranium application (Shin et al., 2002). The presence of Doyurum, 2008), graphene oxide (Chen et al., 2013) and orange peels
U(VI) in wastewater is dangerous to human even in relatively low con- (Mohamoud, 2013). Recently, attention has been focused on the use
centration. The tolerance limit for the discharge of U(VI) into waters is of nanoparticle minerals as adsorbents. Nanoparticle minerals have
50 mg/L (Shuibo et al., 2009). The removal of U(VI) from industrial efflu- better efficiency in removing metal ion pollutants from aqueous solu-
ents is important before discharging them into aquatic environments tions, mainly due to their high surface area to volume ratio (Savage
or onto land. Several techniques of physical and chemical processes and Diallo, 2005; Sharma et al., 2009). The use of alumina nanopar-
have been used for U(VI) removal from industrial wastewaters and ticles for the removal of heavy metals has been investigated due to
radioactive wastes, such as; chemical precipitation, solvent extraction, its low cost, high surface area, surface reactivity, porosity, mechanical
membrane processes, cementation, electrochemical, evaporation and strength, and thermal stability (Stietiya and Jim, 2014). The objective of
adsorption process (Qadeer and Hanif, 1994). Many of these methods the present work was to study the feasibility of using aluminum oxide
are expensive and time consuming and have some limitations. Adsorp- nanoparticles for preconcentration and recovery of U(VI) from aqueous
tion technique is a simple and cost-effective technology that has been solutions. The adsorption kinetics, isotherms and thermodynamics of


Correspondence to: Chemical Engineering Department, Faculty of Engineering, Jazan University, P.O. Box 114, Jazan, Saudi Arabia.
Tel.: +966 564442596.
E-mail address: DrChemEng@yahoo.com
http://dx.doi.org/10.1016/j.psep.2016.02.008
0957-5820/© 2016 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53 45

The adsorption selectivity of AONP for U(VI) ion was carried


Table 1 – Characteristics of aluminum oxide
nanopowder (AONP). out using a multi-metal solution containing Pb(II), Cd(II), Zn(II),
Ni(II) prepared from metal salts. The inlet solution consisted
Parameters Value
of 25 mg/L U(VI) and 10 mg/L of other interfering ions. The
Purity 99.9% adsorption was carried out at the optimum conditions. The
Particle size <20 nm adsorption selectivity coefficient (SC) is defined as follows
Appearance White powder
(Chen et al., 2011):
Nanoparticles (Al2 O3 ) morphology Nearly spherical
Density 3.97 g cm−3  
BET surface area >138 m2 g−1 (qe /Ce )uranium
SC = log (3)
(qe /Ce )other metals

U(VI) ions were studied in batch experiments. Desorption and recovery Desorption of U(VI) from loading AONP was carried out
of uranium ions are also studied.
using different concentrations of HCl, HNO3 and H2 SO4 solu-
tion. Then U(VI) ion in the eluent solution was measured and
2. Materials and methods desorption percent (De%) was calculated from to the following
equation:
2.1. Preparation of adsorbent material
Ce
De% = × 100 (4)
Aluminum oxide nanopowder (␥-nano-Al2 O3 ) (AONP) used in C0
this study was obtained from Beijing, China, Mainland (Model
Number: DK420). The characteristics of Al2 O3 nanopowder 2.4. Adsorption isotherms study
provided by the supplier are presented in Table 1.
The X-ray diffraction analysis (XRD) and environmental Equilibrium studies are described by Freundlich, Langmuir
scanning electron microscope (ESEM) of AONP is also per- and Dubinin–Radushkevich (D–R) isotherms, which are impor-
formed before and after adsorption. tant to explain the relation between the quantity of adsorbed
ions and that remained in the solution at equilibrium.
2.2. Preparation of uranium(VI) stock solution
2.4.1. Langmuir isotherm
Stock solution (1000 mg/L) of uranium(VI) was prepared from The Langmuir equation is used to estimate the maximum
UO2 (NO3 )2 ·6H2 O, Sigma Aldrich, USA. For experiments, the adsorption capacity corresponding to complete monolayer
required concentrations were prepared by dilution. The pH of coverage on the adsorbent surface and is expressed by
solutions was adjusted by 0.1 M Na2 CO3 or 0.1 N HCl. All chem- (Langmuir, 1918):
icals and reagents used in this work, including HCl, HNO3 ,
qmax KL Ce
H2 SO4 and Na2 CO3 were analytical reagent grade. Distilled qe = (5)
1 + KL Ce
water was used throughout the experiments. The other stock
solutions of various metals (Pb(II), Cd(II), Zn(II), Ni(II)) used as
where qe and Ce are the amount of U(VI) adsorbed per gram
interfering ions were prepared as chloride salts.
of adsorbent (mg/g) and the U(VI) concentration (mg/L) in the
aqueous solution at equilibrium, respectively. qmax and KL are
2.3. Adsorption and desorption constants related to the maximum adsorption capacity (mg/g)
and the intensity of adsorption (L/mg), respectively. The linear
Batch experiments were carried out to determine the opti- form of the above equation after rearrangement is given by:
mum parameters of the adsorption process (contact time, pH,
initial concentration, adsorbent dose and temperature). 50 ml Ce 1 Ce
= + (6)
of 25 mg/L of uranium(VI) solutions with a range of pH val- qe qmax KL qmax
ues from 3 to 10 was transferred in a conical glass flask with
100 mg of AONP. The mixture was agitated at 200 rpm in a The constants qmax and KL can be determined from the
thermostatic shaker water bath for different time intervals at slope and intercept of plotting Ce /qe against Ce , respectively.
different temperatures (303, 313, 323 and 333 K). The samples
were withdrawn and centrifuged at 5000 rpm for 5 min and the 2.4.2. Freundlich isotherm
supernatant solutions were analyzed spectrophotometrically Freundlich model (Denise et al., 2012), is used to estimate the
using Shimadzu UV-VIS-1601 spectrophotometer. Adsorption adsorption intensity of the adsorbent toward the adsorbate
capacity (q) of U(VI) was defined as: and is given by:

(C0 − Ce )V qe = KF Ce
(1/n)
(7)
q= (1)
M
where KF is the constant of Freundlich model and n gives an
where C0 and Ce are the initial and equilibrium concentration
indication about the favorability of the adsorption process.
of U(VI) in the aqueous solution (mg/L), respectively. V and
The above equation is conveniently used in linear form as:
M are the volume of solution (L) and mass of adsorbent (g),
respectively. In addition, the removal efficiency (Re ) is calcu- 1
lated according to the following equation: ln qe = ln KF + ln Ce (8)
n
C − C 
0 e A plot of ln Ce against ln qe yielding a straight line indicates
Re (%) = × 100 (2)
C0 the conformation of the Freundlich adsorption isotherm. The
46 Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53

Fig. 1 – ESEM scanning of unloaded (a) and loaded (b) AONP.

constants 1/n and ln KF can be determined from the slope and 2.5.1. Pseudo-first-order kinetic model
intercept, respectively. The pseudo first-order model is assumes that one adsorbate
is adsorbed onto one adsorption site on the adsorbent surface
2.4.3. The Dubinin–Radushkevich (D–R) isotherm (Lagergren, 1898). The linear form of pseudo first order model
The Dubinin–Radushkevich isotherm model was used to pre- was given by equation:
dict the nature of adsorption processes as physical or chemical
by calculating sorption energy (Nguyen and Do, 2001). k1
The linear form of this model is described as: log(qe − qt ) = log qe − t (12)
2.303

ln qe = ln Q0 − KDR ε2 (9) where qe and qt are the amount of U(VI) ions adsorbed onto
AONP at equilibrium (mg/g) and at time t, respectively. k1
where qe is the amount of U(VI) adsorbed onto AONP surface is the adsorption rate constant of pseudo-first-order model
(mol/g) and Q0 represents the maximum adsorption capacity (L/min). Values of k1 and qe were calculated from the slope
of nano-Al2 O3 (mol/g), KDR is a constant related to the mean and intercept of the straight line of plotting log (qe − qt ) versus
adsorption energy (mol2 /kJ2 ) and ε is the Polanyi potential, t, respectively.
which can be calculated from the following equation:
 1
 2.5.2. Pseudo-second-order kinetic model
ε = RT ln 1 + (10) The linear form of pseudo-second order model (Ho and McKay,
Ce
1998) is given by the equation:
The slope of the plot of ln qe versus ε2 gives KDR (mol2 /kJ2 )
and the intercept yields the adsorption capacity Q0 (mg/g). T t 1 1
= + t (13)
is the absolute temperature in K and R is the universal gas qt k2 q2e qe
constant (8.314 J/mol K). The constant KDR gives an idea about
the mean adsorption energy (E) (kJ/mol) of adsorption per mole where k2 is the adsorption rate constant of pseudo-second-
of the adsorbate using the following relationship: order (g/mg/min). The plot of t/qt versus t should give a straight
line and the k2 and qe were calculated from the values of inter-
1
E= √ (11) cept and slope, respectively.
2KDR

2.5. Adsorption kinetics 2.5.3. The Elovich kinetic model


The Elovich model equation is generally expressed as (El-
Pseudo-first order, pseudo-second order, and Elovich kinetic Halwany, 2013):
models were used to investigate the mechanism of U(VI)
adsorption onto AONP. The comparison between the kinetic dqt
models was evaluated by the correlation coefficients (r2 ). = ˛ exp(−ˇdt) (14)
dt
Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53 47

Table 2 – Parameters of Langmuir, Freundlich and


Dubinin–Radushkevich isotherms at different
temperatures (pH 5, time 40 min, dose 100 mg and
25 mg/L U(VI) solution).
Isotherms Temperature (K)

301 313 323 333

Langmuir isotherm
qmax (mg/g) 12.93 13.55 14.88 15.75
KL (L/mg) 0.773 0.713 0.655 0.478
r2 0.989 0.998 0.999 0.993

Freundlich isotherm
KF (mg/g) 8.58 8.77 10.46 11.43
n 6.233 4.513 3.163 2.161
r2 0.897 0.935 0.889 0.954
Fig. 2 – XRD analysis of raw and loaded AONP. D–R model isotherm
Q0 (mg/g) 10.450 10.719 10.95 11.01
KDR (mol2 /kJ2 × 10−5 ) 0.001 0.011 0.042 0.102
E (kJ/mol) 4.071 3.162 1.118 0.873
where ˛ is the initial adsorption rate (mg/g min−1 ), and the
r2 0.933 0.911 0.875 0.942
parameter ˇ is related to the extent of surface coverage and
activation energy for chemisorption (g/mg). The linear form
of Elovich equation is given by the equation: 3.3. Adsorption kinetics

1 1 3.3.1. Effect of pH and initial U(VI) concentrations


qt = ln(˛ˇ) + ln(t) (15) Adsorption of U(VI) onto AONP was found to increase with
ˇ ˇ
increasing time and attained a maximum adsorption capac-
ity attained at 40 min (Fig. 4a). The uptake is studied with
A plot of qt versus ln (t) should yield a linear relationship
changing pH in the range 3–10 at 25 mg/L initial U(VI) concen-
with a slope of (1/ˇ) and an intercept of (1/ˇ) ln (˛ˇ).
tration, 303 K, and adsorbent dose 100 mg. Fig. 4a shows that
the adsorption capacity was highly dependent on pH because
3. Results and discussion the pH of the solution influences not only the distribution of
active sites on the surface of AONP, but also influences the
3.1. Characteristics of adsorbent configuration of U(VI) species. The competition of H+ with
U(VI) on the adsorption sites leads to decrease the removal
Environmental scanning electron microscope (ESEM) of AONP percent of U(VI) at low pH and the increase in U(VI) uptake
is performed before and after adsorption to see the presence of with increasing pH may be due to the decrease in compe-
U(VI) on the surface of AONP which is shown in Fig. 1. Fig. 1(b) tition between H+ and U(VI) ions and consequently more
shows the presence of U(VI) on the surface of nanopowder negatively charged sites were formed (Yusan and Doyurum,
which indicates that AONP has adsorption ability for removal 2008). Above pH 5, the removal efficiency decreases as pH
of U(VI) ions from aqueous solutions. According to the data increases, this was due to the formation of stable complexes
of the joint committee on powder diffraction standards, the UO2 CO3 , [UO2 CO3 ]2− (Mohamoud, 2013). On changing the ini-
XRD analysis of AONP before and after adsorption in Fig. 2 tial U(VI) concentration from 25 to 150 mg/L at 40 min, 303 K
showed no change in the diffraction peaks of AONP before and and pH 5, the adsorption capacity increased from 12.35 to
after the uptake of U(VI) ions. This result indicates that the 49.42 mg/g (Fig. 4b). The results indicate that the optimum
interaction between U(VI) ions and adsorption sites of AONP pH was 5 and the adsorption reaction can be approximated
may be physical in nature. with the pseudo second-order kinetic model (Fig. 5), with
higher values of the correlation coefficient. The rate constants
3.2. Adsorption isotherm and values of correlation coefficient are represented in
Table 3.
Adsorption data of isotherms (Table 2) indicate that the
Langmuir isotherm appears to be the best fitting model for 3.3.2. Effect of adsorbent dose and temperature
adsorption of U(VI) onto AONP because it gives the highest In order to attain the optimal amount of AONP for the adsorp-
correlation coefficient value than other isotherms, indicating tion of U(VI) ions, the adsorption experiments were carried
that monolayer surface adsorption may exist (Fig. 3). To esti- out by adding 50–250 mg of AONP to a series of 50 ml 25 mg/L
mate the mean free energy of adsorption process (E), the rate U(VI) ion solutions at pH 5, 25 mg/L initial U(VI) concentra-
constant (KDR ) of Dubinin and Radushkevich isotherm is used tion, contact time 40 min and 303 K. When the amount of
in Eq. (11). If the value of E = 8–16 kJ/mol then the adsorption AONP exceeded 150 mg, the adsorption percentage of U(VI) on
process flows by chemical ion-exchange, and if E < 8 kJ/mol the AONP was close to 100% due to increase in the adsorption sites
adsorption process is physical in nature, whereas if the value (Mohamoud, 2013) (Fig. 4c). Therefore, 150 mg of AONP was the
is more than 16 kJ/mol, the adsorption process is chemisorp- optimum amount for the adsorption of U(VI) ions. An increase
tion in nature (Nguyen and Do, 2001). The results present in in temperature from 303 K to 333 K lead to increase the adsorp-
Table 2 showed that E value was lower than 8 kJ/mol, indicating tion capacity and a decrease in the equilibrium time to 20 min
that the adsorption of U(VI) onto AONP is physical adsorption at 333 K. This is because increasing the temperature lead to
in nature. increase the adsorption sites of AONP. Hence that the uptake
48 Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53

Fig. 3 – Langmuir (a), Freundlich (b) and Dubinin–Radushkevich (c) isotherms at different temperatures.

Fig. 4 – Adsorption dynamics of U(VI) ions onto AONP at different pH (a) and initial U(VI) concentrations (b) and adsorbent
doses.

Fig. 5 – Pseudo-second-order kinetic model of U(VI) adsorption onto AONP at different pH (a) and initial U(VI) concentrations
(b).

of U(VI) ions onto AONP was found to be fast, and equilibrium of the experiments of U(VI) Pb(II), U(VI) Zn(II), U(VI) Cd(II),
time was attained in 20 min with increasing temperature from U(VI) Ni(II) and U(VI) [Pb(II) Cd(II) Zn(II) Ni(II)] systems.
303 K to 333 K. The results obtained for each of the kinetic The results indicated that AONP has excellent adsorption
models show a better compliance with the pseudo second- selectivity for U(VI) ions from aqueous solution. These results
order equation, with values of correlation coefficient, r2 ∼
= 0.99 suggest that AONP can probably be used in the removal of U(VI)
(Fig. 6). from multi-ionic aqueous system.

3.3.3. Effect of interfering ions 3.4. Mechanism of adsorption


A number of experiments were carried out to study the effect
of interfering ions in the adsorption capacity of AONP for U(VI) In order to investigate the mechanism of adsorption of
ions at pH 5, 150 mg dose, contact time 40 min, initial con- U(VI) ions onto AONP, the obtained kinetic data were eval-
centration of 25 mg/L and 303 K. Table 4, shows the results uated using Boyd’s kinetic model (Riaz, 2007) and Webber’s
Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53 49

Table 3 – Parameters of kinetic models at different temperatures (pH 5, time 40 min, absorbent dose 100 mg and 25 mg/L
U(VI) solution).
Kinetic models Parameter Temperature (K)

303 313 323 333

Pseudo-first-order model k1 (1/min) 0.0032 0.0064 0.0077 0.0095


qe (mg/g) 9.34 9.753 10.135 10.410
r2 0.823 0.875 0.758 0.910

Pseudo-second-order qe (mg/g) 12.23 12.73 13.76 14.01


model
k2 (g/mg/min) 0.0137 0.0134 0.0127 0.010
r2 0.999 0.997 0.989 0.996

Elovich model ˇ (g/mg) 2.033 2.227 2.965 3.121


˛ (mg/g min−1 ) 4.435 5.015 5.366 7.231
r2 0.921 0.874 0.932 0.926

Fig. 6 – Pseudo-second-order kinetic model of U(VI) adsorption onto AONP at different adsorbent doses (at) and
temperatures (b).

Table 4 – Adsorption selectivity of AONP for U(VI) from interfering ions systems.
System Adsorption Adsorption Removal
capacity (mg/g) selectivity efficiency (%)

U(VI) Pb(II) 6.248 ∞ 99.98


U(VI) Zn(II) 6.245 5.79 99.92
U(VI) Cd(II) 6.246 ∞ 99.97
U(VI) Ni(II) 6.248 5.99 99.98
U(VI) [Pb(II) Cd(II) Zn(II) Ni(II)] 6.171 4.89 98.73

intraparticle model (Venkat et al., 2007). Boyd’s kinetic model expression was modified to distinguish between particle diffu-
is expressed as: sion and film diffusion by the following equations (Riaz, 2007):
For intra-particle diffusion:
6  1

F =1− exp(−m2 Bt) (16)


2 m2 Bt = −0.4977 − ln(1 − F) (19)
m=1

where m is an integer that defines the infinite series solution For film diffusion:
and F is the fractional attainment of equilibrium at time t and
is obtained by using the following equation Rt = −ln(1 − F) (20)

qt
F= (17) where R is the rate constant for film diffusion and the values
qe
of Bt and Rt can be determined from values of F. The linear
plots of Bt and Rt versus t according to Eqs. (19) and (20), are
where qt and qe are the amounts adsorbed at time t, and at
shown in Fig. 7 at various temperatures, give an idea about the
equilibrium, respectively, and B is the Boyd’s constant (min−1 )
type of the rate-controlling step involved during the progress
and is obtained by using the following equation:
of adsorption process. Fig. 7a shows the linear plot of Rt versus
t which do no pass through the origin at various temperatures,
Di m2
B= (18) this indicate that the film diffusion is not involved in the kinet-
r2
ics of U(VI) adsorption on AONP. Likewise, Fig. 7b depicts the
where Di is the effective diffusion coefficient of the adsor- linear plot of Bt versus t, which are passed through the origin at
bate in the adsorbent phase and r is the radius of adsorbent various temperatures. This shows that intra-particle diffusion
particles which is assumed to be spherical. Boyd’s model is the rate-controlling step in the mechanism of adsorption
50 Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53

Fig. 7 – Plots of Rt (a) and Bt (b) versus t and Morris–Weber plots for the adsorption of U(VI) ions on AONP at different
temperatures.

Fig. 8 – Van’T Hoff plot of thermodynamic parameters (a) plot of ln k2 versus 1/T for activation energy (b) of U(VI) adsorption
onto AONP nanopowder.

of U(VI) ions onto AONP. A further study was carried out by of temperature using the following equations (El-Halwany,
Weber–Morris kinetic model to confirm the rate limiting step 2013):
of adsorption process by the following equation (Venkat et al.,
2007): S◦ H◦
ln KD = − (22)
R RT

q
qt = Kid t0.5 (21) KD = (23)
Ce

G◦ = −RT ln KD (24)


where Kid and t are the rate constant of intraparticle diffusion
(mg/g min−1/2 ) and contact time (min), respectively. The value The values of H◦ , S◦ were determined from the slope
of Kid is evaluated from the slope of the straight line of qt versus and intercept of the straight line of plotting ln KD versus 1/T,
t0.5 . When intraparticle diffusion plays a role in controlling the respectively (Fig. 8a). The values of H◦ , S◦ and G◦ were
kinetics of the sorption process, the plots of qt versus t0.5 yield given in Table 5. The positive value of H◦ indicated the
straight lines passing through the origin and the slope gives endothermic nature of adsorption process and the negative
the rate constant kid . From Fig. 7c, It is also observed that the value of G◦ showed the spontaneous nature of the uptake
plots of qt versus t0.5 yield straight lines passing through the of U(VI) onto AONP. The magnitude of H◦ may give an addi-
origin at various temperatures, confirming that the intraparti- tional idea about the type of adsorption process. If the value
cle diffusion is the rate-controlling step in the mechanism of of H◦ = 2.1–20.9 kJ/mol then the adsorption process proceeds
adsorption of U(VI) ions onto AONP. by physical adsorption, and if H◦ = 80–200 kJ/mol the adsorp-
tion process is chemisorption in nature (El-Halwany, 2013).
From Table 5, the value of H◦ (17.91 kJ/mol) indicates that the
3.5. Thermodynamic studies adsorption of U(VI) onto AONP is physical adsorption in nature.

According to Van’t Hoff equation, thermodynamic parameters, 3.6. Determination of activation energy
such as enthalpy variation (H◦ ), entropy variation (S◦ ) and
change in Gibbs free energy (G◦ ), were calculated from the The pseudo-second-order model is identified as the best
curve relating the distribution coefficient (KD ) as a function kinetic model for the adsorption of U(VI) onto AONP.
Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53 51

Table 5 – Thermodynamic parameters of U(VI) adsorption onto AONP (pH5, time 40 min, absorbent dose 100 mg and
25 mg/L U(VI) solution).
Temperature (K) ln KD H◦ (kJ/mol) S◦ (kJ/mol/K) G◦ (kJ/mol)

303 3.71 −17.91 −0.0892 −9.346


313 3.80 −9.888
323 3.92 −10.52
333 4.52 −12.51

of the desorption process, uranium ions can be recovered from


Table 6 – Desorption efficiency of eluting agents at
different concentrations. eluting solution by precipitation process using NH4 OH (33%) at
pH 8 to produce a precipitate of ammonium di-uranate (yel-
Eluting Desorption efficiency (%)
low cake) (ADU) ((NH4 )2 U2 O7 ) (Man et al., 2002; Goswami and
agents
Concentration of eluting agents Das, 2003). The precipitate was separated by filtration using
ashless paper. The precipitate (yellow cake) was sintered in
0.5 M 1M 1.5 M 2M 2.5 M 3M an electric furnace at 1073 K to produce uranium oxide U3 O8
(Mahmoud, 2011). After cooling the solid sample was analysis
HCl 30.32 45.21 57.89 64.86 64.85 64.86
HNO3 65.45 93.43 99.98 99.99 99.99 99.99
and the results are shown in Fig. 9. This indicates that AONP is
H2 SO4 44.56 56.54 63.54 70.65 70.64 70.59 a good adsorbent material for U(VI) ions and could be success-
fully applied for preconcentration and recovery of U(VI) ions
from aqueous solution.
Table 7 – Adsorption and desorption efficiency during
five adsorption–desorption cycles in the batch technique.
3.8. Batch adsorber design
Number of cycle Adsorption (%) Desorption (%)

1 99.85 99.99 From adsorption isotherm studies the best fit of adsorption


2 99.98 99.88 isotherm was used to design a single unit of batch adsorption
3 99.79 99.89 process (Mahmoud, 2011; Priya and Venkateswaran, 2014). A
4 99.99 99.56 schematic diagram of single-stage of batch adsorption process
5 98.87 99.99
was shown in Fig. 10a. Consider an effluent containing V (m3 )
of liquid waste and let the U(VI) concentration got reduced
Accordingly, the rate constant k2 of the pseudo-second-order from C0 to C1 (g U(VI)/m3 liquid waste). The amount of adsor-
model is used to calculate the activation energy (Ea ) of the bent was M (kg) and the U(VI) loading changed from q0 to q1
adsorption process using the Arrhenius equation (El-Halwany, (g U(VI) per kg adsorbent). When fresh adsorbent is used, q0 = 0
2013). and the mass balance equates the U(VI) removed from the
liquid waste to that picked up by the adsorbent. The mass
ln A0 − Ea
ln k2 = (25) balance equation for the adsorption system in Fig. 10a, can be
RT
written as
where k2 , A, Ea , R, and T are the rate constant of the pseudo-
second-order model (g/mg min), Arrhenius factor, activation V(C0 − C1 ) = M(q0 − q1 ) = Mq1 (26)
energy (kJ/mol), gas constant (8.314 J/mol K) and temperature
M C0 − C1
(K), respectively. The activation energy can be determined = (27)
V q1
from the slope of the plot of ln k2 versus 1/T (Fig. 8b). The
magnitude of activation energy may give another indication
At equilibrium, C1 = Ce and q1 = qe . So that Eq. (27) can be
about the type of U(VI) adsorption onto AONP. The physisorp-
written as:
tion process usually has the activation energy (Ea ) in the range
of 5–40 kJ/mol, while the activation energy (Ea ) in the range of M C0 − Ce
= (28)
40–800 kJ/mol suggests the chemisorption process (Mahmoud V qe
and El-Halwany, 2014; Mahmoud, 2014). In this study, the acti-
vation energy (Ea ) was 6.94 kJ/mol, indicating that the process In the case of the adsorption of U(VI) onto AONP, it had been
is physisorption. This result also conforms to the conclusion already seen that the Langmuir isotherm gave the best fit to
derived from the above results from the mean free energy of experimental data. Consequently, the Langmuir equation can
adsorption (E) and value of H◦ . be best substituted for qe in the rearranged form of Eq. (28) as
follows:
3.7. Desorption and recovery of U(VI) ions M C0 − Ce C0 − Ce
= = (29)
V qe KL Ce /(1 + qmax Ce )
Desorption efficiency and reusing of AONP were investigated
in batch method using various concentrations of HCl, HNO3 The calculated design of a single stage of batch
and H2 SO4 solutions at pH 5, contact time 40 min and 303 K. adsorber for treating 60 m3 of different initial U(VI)
Table 6, shows that 1.5 M HNO3 solution can effectively des- concentration(50–400 mg/L) was used to calculate the amount
orb U(VI) ions up to 100% from the loaded adsorbent material. of AONP required for the removal of U(VI) to the extent of
AONP was reused and adsorption–desorption cycles were car- 99.85% at optimum condition (Fig. 10b). The results of present
ried out for 5 cycles. The results in Table 7, showed that the investigation showed that aluminum oxide nanopowder have
adsorption percent of AONP was not significantly changed as considerable potential for the removal of the U(VI) ions from
the adsorption–desorption cycles proceeded. On completion liquid waste over a wide range of concentrations.
52 Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53

Fig. 9 – ESEM scanning of sintering ammonium di-uranate (yellow cake) at 1073 K.

Fig. 10 – Single-stage of batch adsorption process (a) and ANOP (M) required for treated 60 m3 volumes to 98.86% removal at
different initial U(VI) concentration 50–400 g/m3 .

4. Conclusion these results AONP displays a good capacity and an excellent


selectivity for preconcentration and recovery of uranium ions
The results of this work indicate that AONP is an excellent from an aqueous solution. A single unit of batch adsorber has
adsorbent material has a good selectivity for U(VI) adsorp- been designed for various initial U(VI) concentration of liquid
tion from aqueous solution. The results suggested that the wastes using the Langmuir isotherm.
adsorption process was dependent on uptake time, initial
U(VI) ion concentration, pH, adsorbent dose and temperature.
Adsorption data indicates that the process is following Lang- References
muir isotherm and pseudo-second-order kinetic model. The
Abbasi, W.A., Streat, M., 1994. Adsorption of uranium from
rate-controlling step in the process appeared to be influenced
aqueous solutions using activated carbon. Sep. Sci. Technol.
by intraparticle diffusion. Thermodynamic parameters indi-
29, 1217–1230.
cate that adsorption processes is endothermic in nature and Chen, S., Hong, J., Yang, H., Yang, J., 2013. Adsorption of
the rate of U(VI) uptake was found to be faster with increas- uranium(VI) from aqueous solution using a novel graphene
ing temperature. The values of mean free energy obtained oxide-activated carbon felt composite. J. Environ. Radioact.
from Dubinin and Radushkevich isotherm (E) indicate that the 126, 253–258.
process is physisorption, which is confirming the conclusion Chen, X.Q., Koon, F.L., Shuk, F.M., King, L.Y., 2011. Precious metal
recovery by selective adsorption using biosorbents. J. Hazard.
derived from activation energy (Ea ) and from the value of H◦ .
Mater. 186, 902–910.
Desorption of uranium ions can be carried out using 1.5 M
Denise, A., Fungaro, I., Yamaura, M., Craesmeyer, R.G., 2012.
HNO3 and uranium ion can be recovered by precipitation. From Uranium removal from aqueous solution by zeolite from fly
Process Safety and Environmental Protection 1 0 2 ( 2 0 1 6 ) 44–53 53

ash-iron oxide magnetic nanocomposite. Int. Rev. Chem. Eng. Priya, V.T., Venkateswaran, V., 2014. Equilibrium and kinetic data
4, 353–358. and process design of methylene blue onto stishovite-TiO2
El-Halwany, M.M., 2013. Kinetics and thermodynamics of nanocomposite. Glob. J. Biol. Agric. Health Sci. 3,
activated sunflowers seeds shell carbon (SSSC) as sorbent 102–111.
material. J. Chromatogr. Sep. Tech. 4, 5–11. Qadeer, R., Hanif, J., 1994. Kinetics of uranium(VI) ions adsorption
Goswami, D., Das, A.K., 2003. Preconcentration and recovery of on activated charcoal from aqueous solutions. Radiochim.
uranium and thorium from Indian monazite sand by using a Acta 65, 259–263.
modified fly ash bed. J. Radioanal. Nucl. Chem. 258, 249–254. Riaz, R., 2007. Adsorption behavior of ruthenium ions on
Ho, Y.S., McKay, G., 1998. The kinetics of sorption of basic dyes activated charcoal from nitric acid medium. Colloids Surf. A
from aqueous solution by sphagnum moss peat. Can. J. Chem. 293, 217–223.
Eng. 76, 822–827. Savage, N., Diallo, M.S., 2005. Nanomaterials and water
Konstantin, M., Demetriou, A., Pashalidis, I., 2007. Adsorption of purification: opportunities and challenges. J. Nanopart. Res. 7,
hexavalent uranium on dunite. Glob. NEST J. 9, 229–236. 331–342.
Lagergren, S., 1898. About the theory of so-called adsorption of Sharma, Y.C., Srivastava, V., Singh, V.K., Kaul, S.N., Weng, C.H.,
solution substances. Handlinge 24, 147–156. 2009. Nanoadsorbents for the removal of metallic pollutants
Langmuir, I., 1918. The adsorption of gases on plane surfaces of from water and wastewater. Environ. Technol. 30,
glass, mica and platinum. J. Am. Chem. Soc. 40, 1361–1403. 583–609.
Mahmoud, M.A., El-Halwany, M.M., 2014. Adsorption of cadmium Shin, D.C., Kim, Y.S., Moon, J.Y., Park, H.S., Kim, J.Y., Park, S.K.,
onto orange peels: isotherms, kinetics, and thermodynamics. 2002. International trends in risk management of
J. Chromatogr. Sep. Tech. 5, 238–243. groundwater radionuclides. J. Environ. Toxicol. 17,
Mahmoud, M.A., 2014. Thermodynamic and kinetic studies: 273–284.
adsorption of Fe(III) onto corncob. J. Thermodyn. Catal. 5, 1–6. Shuibo, X., Chun, Z., Xinghuo, Z., Jing, Y., Xiaojian, Z., Jingsong,
Mahmoud, M.A., (Ph.D. thesis) 2011. Preconcentration and W., 2009. Removal of uranium(VI) from aqueous solution by
recovery of uranium from radioactive liquid wastes. Faculty of adsorption of hematite. J. Environ. Radioact. 100, 162–166.
Engineering, Elmina University, Egypt. Stietiya, M.H., Jim, J.W., 2014. Zinc and cadmium adsorption to
Mohamoud, M.A., 2013. Removal of uranium(VI) from aqueous aluminum oxide nanoparticles affected by naturally occurring
solution using low cost and eco-friendly adsorbents. J. Chem. ligands. J. Environ. Qual. 43, 498–506.
Eng. Process Technol. 4, 200–206. Venkat, S., Mane, I.D.M., Vimal, C.S., 2007. Kinetic and
Man, J.K., Bo, E.H., Pil, S.H., 2002. Precipitation and adsorption of equilibrium isotherm studies for the adsorptive removal of
uranium(VI) under various aqueous conditions. Environ. Eng. Brilliant Green dye from aqueous solution by rice huskash. J.
Res. 7, 149–157. Environ. Manag. 84, 390–400.
Nguyen, C., Do, D.D., 2001. The Dubinin–Radushkevich equation Yusan, S., Doyurum, A.S., 2008. Sorption of uranium(VI) from
and the underlying microscopic adsorption description. aqueous solutions by akaganeite. J. Hazard. Mater. 160,
Carbon 39, 1327–1336. 388–395.

You might also like