Synthesis, characterization and application of ... - SAGE Journals

3 downloads 0 Views 731KB Size Report
The flasks were placed on a shaker at different temperatures (25–60. C), and ..... El Sayed MS (2003) Uranium extraction from gattar sulfate leach liquor using ...
Research Article

Synthesis, characterization and application of composite derived from rice husk ash with aluminium oxide for sorption of uranium W. M. Youssef, M. S. Hagag

Adsorption Science & Technology 0(0) 1–20 ! The Author(s) 2018 DOI: 10.1177/0263617418768920 journals.sagepub.com/home/adt

and A. H. Ali

Nuclear Materials Authority, Egypt

Abstract A composite of rice husk (RH), caustic soda and aluminium oxide was synthesized at 500 C. The activated carbon and amorphous silica dispersed over the aluminium oxide selectively adsorbed uranium in the presence of other elements. At equilibrium time 1 h, phase ratio S/L (0.1 g/10 ml), pH ¼ 5 and uranium initial concentration 120.6 mg/l uranium adsorption efficiency was 96.35%. The uranium stripping efficiency from the load RHA–alumina composite fulfilled 99.9% at 1 h equilibrium time, a phase ratio (S/A) of 0.05 g/10 ml and 0.5 mol/l HNO3. The scanning electron microscopy photos revealed that the rice husk ash (RHA)–alumina composite has vacant or regular cavities before the adsorption, and the cavities are fully occupied by uranium after adsorption. The Fourier transform infrared spectroscopy shows a more broadening of the band t ¼ 3526 and 3462 cm1 which was ascribed to the uranium adsorption. The composite adsorbed 93.75% of uranium from a waste sample. The uranium adsorption exhibited a Langmuir isotherm. Keywords Rice husk ash, uranium adsorption, activated carbon, amorphous silica Submission date: 22 October 2017; Acceptance date: 23 February 2018

Corresponding author: M. S. Hagag, Nuclear Materials Authority, 530 P.O Box Maadi, Cairo, Egypt. Email: [email protected] Creative Commons CC BY: This article is distributed under the terms of the Creative Commons Attribution 4.0 License (http://www.creativecommons.org/licenses/by/4.0/) which permits any use, reproduction and distribution of the work without further permission provided the original work is attributed as specified on the SAGE and Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).

2

Adsorption Science & Technology 0(0)

Introduction Rice husk (RH) is a by-product of rice cultivation and harvesting that causes hazardous environmental problems and pollution in Egypt. The cultivated rice region in Egypt encompasses approximately about 810,000 hectares. The Egyptian production of rice is nearly 4.5 million tons per year, according to the Egyptian Ministry of Agriculture. Thus, immense quantities of RH yielded from rice cultivation must be utilized and innovative chemical methods for elimination are needed. RH can be exploited for many applications including transition metal adsorption (Franco et al., 2017), refractory ceramic developments (Sobrosa et al., 2017), cement (Sinyoung et al., 2017), removing chemical impurities (Lee et al., 2017) and catalysts (Setthaya et al., 2017; Shi et al., 2017) and catalysis for biodiesel production (Guo et al., 2017). Two goals must be fulfilled: eliminating the pollutants that are associated with rice cultivation in Egypt and using the chemically modified pollutants to remove some of the other environmentally hazardous pollutants such as heavy elements, i.e. dispose of pollutants using chemically modified pollutants. The major contents of RH are silica and organic matter (cellulose) (Mohamed et al., 2015). The synthesis of a composite with a novel matrix from RH and inorganic compounds (alumina and caustic soda) via ignition results in activated carbon (AC) (produced from organic matter which is the active adsorbent). The silica that is converted to amorphous silica after ignition has a high surface area permitting and enhancing the removing of heavy elements. The AC derived from RH is a porous carbonaceous solid material with a large surface area and high porosity (Hu and Srinivasan, 1999) which allows sorption of wastes from gas and liquid phases (Jankowska et al., 1991). The raw material type used and the activation routes influence the pore characteristics and adsorption capability of AC. AC is normally created through carbonization and activation of the RH precursor. These procedures are essential for the porous structure and overabundant pore volume of RH-derived AC. The carbonization process is typically executed at temperatures ranging from 500 to 900 C to eliminate the non-carbon elements present in RH, such as oxygen, hydrogen and nitrogen as volatile gaseous products (Daud et al., 2000). The formation of free interstitial spaces and the remnant carbon atoms combining into aromatic sheets (cross-linked in a random style) lead to the formation of pores. The pores created in carbonized RH can be further developed via an activation process. Activation widens the existing pores by burning off the walls of the adjacent pores and removing the disorganized carbon that blocks the pores in carbonized RH. Physical, chemical and physiochemical (a combination of two preprocesses, i.e. physical and chemical) activation processes can be used to prepare RH-derived AC (Hameed et al., 2007). Chemical activation of RH-derived AC involves carbonization and impregnation in a single step using chemical agents. Uranium is used in the nuclear fuel industry (Hore-Lacy, 2016) and it can be applied in electronic industries, semiconductors (Adamska et al., 2015; Bacci et al., 1989), catalysis (Amrute et al., 2013; Dong et al., 2015) and alloys (Ahn et al., 2016; Ghoshal et al., 2014). Given the importance of uranium, many chemical methods have been utilized for the recovery and separation of uranium such as solvent extraction (Ahn et al., 2016; Dartiguelongue et al., 2016; Zhu et al., 2016; Ghoshal et al., 2014), ionic exchange resins, (Chen et al., 2016; Ogden et al., 2017), adsorption over modified adsorbent (Gajowiak et al., 2013; Grabias et al., 2013), liquid emulsion membrane (Biswas et al., 2012; El Sayed, 2003), chelating resin (Donia et al., 2009; Ilaiyaraja et al., 2017), ion inclusion membranes (Kolev et al., 2013; St John et al., 2012) and AC sorption (Afsari et al., 2012; Yakout et al., 2013).

Youssef et al.

3

In this article, a composite consisting of RH, caustic soda and alumina was synthesized by mixing and igniting at 500 C and the adsorption potential of the composite for uranium was studied. The ignition of RH with alumina and caustic soda resulted in the dispersion of active sites on the adsorbent surface, which was reflected in enhanced, selective uranium sorption in the presence of impurities. The adsorption of uranium by the RHA–alumina composite was elucidated with infrared (IR) spectroscopy and scanning electron microscope (SEM).

Experimental Preparation of RH–alumina composite Typically, 10 g of RH was obtained from a local farm in the Sharkia governorate. Table 1 shows the chemical compositions of RH. The RH was added to 10 g of Al–carbonate and 20 g of NaOH in a porcelain crucible, and the mixture was polymerized at 500 C for 3 h. A black powder was obtained, washed several times with deionized water to neutralize the powder and remove excess chemicals, and desiccated at 120 C to complete dryness to obtain the RHA–alumina composite.

Material characterization The characteristic main structure of the RHA–alumina composite was determined using different analytical techniques such as:

Fourier transform infrared spectroscopy (FTIR) A Thermo Scientific Nicolet IS10, model instrument Germany.

SEM SEM model Philips XL 30 ESEM (25–30 keV accelerating voltage, 1–2 mm beam diameter and 60–120 s counting time). The minimum detectable weight concentration ranged from 0.1 to 1 wt% with a realized precision less than 1%.

Preparation of standard solutions All chemical reagents used were analytical grade. Stock solutions of 1000 mg/l of thorium and copper were prepared from standard stock solutions and the uranium, calcium and iron Table 1. The chemical composition of the studied rice husk sample. Constituent

wt%

Constituent

wt%

N P K Si C Bulk; density kg/m3

0.3 0.35 0.25 11.3 35 0.99

H O S Moisture Lignin Hardness Mohr’s scale

6 4.5 0.05 10.5 30 5.7

4

Adsorption Science & Technology 0(0)

solutions were prepared from uranium acetate, dehydrated calcium chloride and ferric chloride, respectively, by dissolving the specified weight of the salt in deionized water. Uranium was determined in all the different working aqueous solutions using Arsenazo III dye complexation method (Marczenko, 1976). The uranium and Arsenazo III dye complex that formed was measured at a wavelength of approximately 655 nm against proper standard solutions using a double beam Shimadzu, 1401 UV/VIS spectrophotometer (Japan). The copper and iron concentrations were determined using an atomic absorption spectrometer apparatus. The concentration of Th(IV) in the aqueous solution phase was determined spectrophotometrically using Thorn method. The absorption of Th(IV) was measured for Th(IV) at 540 nm.

Batch experiments (adsorption and elution studies) The effects of different factors on the adsorption process were studied. The studied factors included the solution pH, equilibrium time, uranium concentration and solid/liquid ratio. The adsorption experiments were performed by shaking 0.1 g samples of the prepared RHA–alumina composite with 10 ml of a uranium solution (42.4–3500 mg/l), and the pH was changed from 2 to 10. The effect of the presence of co-ions was demonstrated and calculated using different concentrations (120 mg/l) of iron, copper, calcium and thorium. The flasks were placed on a shaker at different temperatures (25–60 C), and aqueous samples were taken at time periods of 15–120 min. After treatment, the solid phases were separated using Whatman filter paper (no. 40), and the uranium concentration in the filtrate was chemically determined. The quantity of uranium loaded on the composite was calculated using equation (1), i.e. by taking the difference between the initial and residual concentrations of uranium in solution and dividing it by the weight of the adsorbent. The removal or adsorption efficiency (Re) was defined as the U sorption percentage relative to the initial concentration (equation (2)) qe ¼ Re ¼

Co  Ce V M

(1)

Co  Ce  100 Co

(2)

where qe (mg/g) is the amount of uranium loaded per unit weight of the adsorbent; Co and Ce are the initial and equilibrium (or at any time) uranium or interfering ions concentration (mg/l), respectively; V is the volume per litre of solution and M is the weight (g) of the adsorbent (RHA–alumina composite). To strip uranium from the RHA–alumina composite adsorbent, a variety of stripping agents were examined, including HCl, H2SO4, Na2CO3, HNO3 and NaCl.

Liquid waste properties One litre of a waste solution with a pH of 1.3 was obtained from an ore processing unit (Inshass) which later adjusted to working optimum pH 5. The concentrations of some cations and anions in the solution are shown in Table 2.

Youssef et al.

5 Table 2. Concentration of some cations and anions of waste sample. Element

ppm

SO42 CO32 Cl UO2 SiO2 Fe Zr Zn La Mn Mg Ce Cu Pb Ca

53 80 16 100 800 1600 158 10 596 230 1540 190 2 270 1000

Figure 1. FTIR spectra of the RHA–alumina composite before and after uranium sorption.

Results and discussion Characterization FTIR. The loading capacity of the RHA–alumina composite depends on the porosity, pore content and functional group reactivity on the surface. IR spectrum was used to qualitatively interpret the functional groups on the RHA–alumina composite (Figure 1).The IR spectra were similar to the spectrum of some AC and silica. The band at t ¼ 3620 cm1

6

Adsorption Science & Technology 0(0)

Figure 2. SEM images of the RHA–alumina composite (a) before and (b) after the adsorption of uranium.

disappeared, which was attributed to the sorption of uranium. The absorption bands at 1022 cm1 belong to Si–O–Si stretching modes which change significantly before and after uranium adsorption of the absorption bands at 3526 and 3462 cm1 became wider, which was also attributed to the sorption of uranium. The change in the absorption bands can be explained as follows, the adsorption of uranium on the surface of the adsorbent leads to change in the energies needed for vibrational motion and angle bending for the bonds before adsorption and after. Surface morphology studies. The SEM images of the RHA–alumina composite macrostructure before and after the sorption of uranium (Figure 2(a) and (b)) show the rudimentary RHA– alumina composite surface that formed and the surface structure is attributed to the volatility and the decomposition of the hydrothermal treatment. After the adsorption, the pores and crevices were occupied by uranium ions (Figure 2(b)).

Controlling factors on the adsorption process Effect of pH. The pH can greatly influence the uranium adsorption efficiency (A%) and uranium uptake (qe) (mg/g). Synthetic uranium solution pH intervals in the range 2–10 were shaken with RHA–alumina composite and the other experimental conditions were S/A phase ratio 0.1 g/10 ml, 134 mg/l U(VI) concentration, 1 h equilibrium time and room temperature 25  2 C. Figure 3 reveals that the efficiency of the uranium adsorption (A%), uranium uptake (qe) increased from 18 to 96% and a maximum value of 13 mg/g, respectively, at pH value 5. At pH 10, uranium uptake decreases to 3.8 mg/g. Thus, pH 5 was used for the synthetic uranium solutions in the next experiments. At a higher acidity, the uranium (VI) uptake decreased due to the increasing Hþ ion concentrations, which caused the surface of the adsorbent to be increasingly more positive. Then, competition occurs between the Hþ ions, which are small and fast, and the uranium species, which are also positive, and adsorption is not favoured. Thus, the presence of a higher concentration of Hþ ions in the reaction mixture caused a reduction in the uranium uptake. In contrast, increasing the pH of the synthetic uranium solutions resulted in the adsorbent surface becoming more negatively charged, which resulted in the more favourable adsorption of positively charged species (Han et al., 2007).

Youssef et al.

7 100

16 adsorpon effeciency,% uranium uptake qe, mg/g

80 70

14 12

60

10

50

8

40

6

30

4

20

2

10 0

uranium uptake qe, mg/g

Uranium adsorpon effeciency, %

90

0 0

2

4

6

8

10

12

pH Figure 3. Effect of pH on the uranium adsorption efficiency (A%) and uptake (mg/g).

90

adsorpon effeciency,% qe, mg/g

100

80 70

80

60 50

60 40 30

40

20 20

uranium uptake qe, mg/g

Uranium adsorpon effeciency, %

120

10 0 0

500

1000

1500

2000

2500

3000

0 3500

uranium concentraon ppm

Figure 4. The effect of the initial uranium concentration on the uranium adsorption efficiency (A%) and uptake (mg/g).

The effect of the initial uranium concentration. The effect of the initial uranium(VI) concentration on the adsorption was studied. The RHA–alumina composite (0.1 g) was shaken with approximately 10 ml of the uranium solutions at different concentrations ranging from 42.4 to 3500 mg/l, and the other parameters were a pH of 5, 1 h shaking time and ambient temperature 25  2 C. The relationship between the different concentrations and the uranium adsorption efficiency (%A) and uranium uptake (qe) is illustrated in Figure 4. The uranium adsorption efficiency decreased with the increasing uranium concentration because the uranyl ions (UO22þ) have a higher mobility in diluted solutions, which permits more interactions between the adsorbent and uranyl ions. However, increasing the uranium concentration over 120.6 mg/l results in increasing competition of the UO22þ ions for the free active sites, which negatively affects the adsorption efficiency, not the burden capacity.

8

Adsorption Science & Technology 0(0)

The amount of adsorbed uranium and uranium uptake, qe (mg/g), increased with the increasing uranium concentration, and the largest adsorption capacity was at a uranium concentration of 1032.35 mg/l. Uranium concentrations above this level did not negatively or positively influence the uranium adsorption or uptake. Therefore, the maximum uranium loading capacity of the RHA–alumina composite is 85 mg/g. Adsorption isotherms. Several frequently used adsorption isotherm models were used to model the isotherm data from the equipoise adsorption of the RHA–alumina composite. Three of these models are Langmuir, Freundlich and Dubinin isotherms. A. Langmuir isotherm: In the Langmuir model, adsorption uniformly occurs on the active sites of the absorbent, and once a sorbate occupies a site, no further sorption can occur at this site. Thus, the Langmuir model is represented by the following equation (Chegrouche et al., 1997; Mellah and Chegrouche, 1997) Ce =qe ¼ 1=bQo þ Ce =Qo

(3)

where Qo and b are the Langmuir constants for the sorption capacity of the saturated monolayer and the sorption equilibrium constant, respectively. A graph of Ce/qe versus Ce will result in a straight line with a slope of (1/Qo) and an intercept of 1/bQo, as seen in Figure 5. The Langmuir specifications given in Table 3 can be used to anticipate the tendency of the sorbate and sorbent using the dimensionless separation factor RL (Bhatnagar and Jain, 2005; Parab et al., 2005) RL ¼ 1=ð1 þ bCo Þ

(4)

The RL value indicates whether the isotherm is irreversible (RL¼0), proper (0 < RL1) (Fan et al., 2011; Ho and McKay, 1999). The values of RL for the sorption of uranium (VI) onto the RHA–alumina composite are presented in Figure 6, and the values indicate that the adsorption of uranium (VI) is higher at higher initial uranium (VI) concentrations and the opposite was true at lower concentrations. 30 25

Ce/qe

20 15 10 5 0 0

1000

2000

3000

Ce, mg/l

Figure 5. Langmuir isotherm for uranium adsorption on the RHA–alumina composite.

9

RL

Youssef et al. 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

500

1000

1500

2000

2500

Co, mg/l

Figure 6. The separation factor, RL, for uranium(VI) adsorbed on the RHA–alumina composite.

B. Freundlich isotherm The Freundlich model postulates that the ratio of the adsorbed solute to the solute concentration is a function of the solution. The empirical model was shown to be consistent with exponential distribution of active centres, characteristic of heterogeneous surfaces. The quantity of adsorbed solute at equipoise qe, the concentration of the uranium in the solution at equilibrium Ce, is represented in the following equations (Chegrouche et al., 1997; Mellah and Chegrouche, 1997) qe ¼ KF Ce 1=n

(5)

This expression can be linearized to obtain Log qe ¼ logKF þ ð1=nÞlog Ce

(6)

where KF and n are the Freundlich constants related to the adsorption capacity and adsorption intensity, respectively. A plot of logqe versus log Ce results in a straight line with a slope of 1/n and an intercept of logKF as shown in Figure 7. Table 3 presents the Freundlich constants. The experimental results indicate that uranium adsorption on the RHA–alumina composite was better fit by a Langmuir isotherm than Freundlich. C. Dubinin–Radushkevich isotherm model The Dubinin–Radushkevich isotherm model is used to denote the adsorption mechanism with the Gaussian energy distribution on a heterogeneous surface. The model has often successfully fitted with high solute activities and the intermediate range of concentration data well. The Dubinin–Radushkevich isotherm equation is linearly represented as follows Inqe ¼ Inqs –Bd ðeÞ2

(7)

e ¼ R T ln ½1 þ 1=Ce 

(8)

where qe (mg/g) is the adsorbed value of the uranyl ion at equilibrium concentration, qs is the theoretical isotherm saturation capacity (mg/g), Bd is the Dubinin–Radushkevich isotherm

10

Adsorption Science & Technology 0(0) 2.5

log qe

2 1.5 1 0.5 0 0

1

2 log Ce

3

4

Figure 7. Freundlich isotherm for uranium adsorption on the RHA–alumina composite. 4.50 4.40 4.30

log qe

4.20 4.10 4.00 3.90 3.80 3.70 3.60 0.00

10.00

20.00

30.00

e2 Figure 8. Dubinin–Radushkevich isotherm for uranium adsorption on the RHA–alumina composite.

constant (mol2/kJ2), e is the polanyl potential, Ce (mg/l) is U(VI) concentration in leach liquor at equilibrium, R (J/mol K) is the gas constant and T (K) is the absolute temperature. A plot of the logarithm of amount adsorbed In qe versus the square of potential energy e2 where qd and Bd are calculated from the slope and intercept of the linear plots. The mean adsorption energy E (J/mol) of adsorption of uranyl ions to the RHA–alumina composite from infinity in the standard solutions can be estimated from the following equation E2 ¼ 1=2 Bd

(9)

If the magnitude of E is between 8 and 16 kJ/mol, the sorption process is supposed to proceed via chemisorption, but if E is less than 8 kJ/mol, the sorption process is physisorption. The mean adsorption energy E (J/mol) of adsorption of uranyl ions to the RHA–alumina composite is calculated to be 4.22 kJ/mol (Dabrowski, 2001; Dubinin, 1960).

Youssef et al.

11

Table 3. Langmuir and Freundlich parameters for uranium adsorption on the RHA–alumina composite. Langmuir model parameters Metal

Adsorbent

Uranium

RHA–alumina composite

Freundlich model parameters

Dubinin–Radushkevich model

Qo (mg/g)

b (l/mg)

R2

1/n

Kf (mg/g)

R2

E

R2

85

0.029

0.99

0.696

1.33

0.952

4.22

0.847

The effect of the equilibrium time The effect of the equilibrium time on the uranium adsorption efficiency (A%) from a 10 ml solution of uranium (120.6 mg/l) using the RHA–alumina composite was studied. The time was varied from 15 to 120 min at ambient temperature (25  2 C) with a pH of 5 and approximately 0.1 g of the RHA–alumina composite. The data are plotted in Figure 9, and a gradual increase in the uranium adsorption efficiency was observed with the increasing equilibrium time. A maximum value of 96% was attained at 60 min, and then the adsorption remained constant. Hence, the adsorption equilibrium time used for the subsequent experiments was 60 min.

The effect of the solid/liquid ratio(S/a) The influence of the solid/liquid ratio, g/ml, was studied in the range varied from 2.5 to 15 g/ml on the adsorption efficiency of uranium (VI) from a solution assaying 120.6 mg U/l onto RHA–alumina composite. The other parameters were fixed on shaking time 60 min, ambient temperature 25 2 C and a pH of 5. The results plotted in Figure 10 show the relation between the uranium adsorption efficiency, A%, and uranium uptake quantity and the solid/liquid ratio. An increase in the solid/liquid ratio from 2.5 to 10 g/l clearly caused an increase in the adsorption efficiency from 66.83 to 96.35%, but the uranium uptake decreased from 32.24 to 11.66 mg/g. Therefore, a liquid/solid ratio, L/g used for the other experiments was 10 g/l.

The effect of the temperature The effect of the temperature on the uranium adsorption efficiency with 0.1 g of the RHA– alumina composite in a 10 ml solution with 120.6 mg/l of standard uranyl nitrate was studied in the temperature range from 25 to 60 C. The other factors were held constant, i.e. a pH of 5 and 60 min shaking time. The results in Figure 11 show that the uranium adsorption efficiency of the RHA–alumina composite decreased with the increasing temperature, which indicated that the U(VI) adsorption was favoured at ambient temperature.

The effect of competing ions The influence of other interfering metal ions, such as copper, thorium and calcium, which may be present along with uranium ions in synthetic aqueous solutions, was studied by separately adding different cations to a uranium solution under the following conditions:

12

Adsorption Science & Technology 0(0)

Uranium adsorption efficiency, (%A)

adsorption effeciency,%

uranium uptake qe, mg/g

100 90 80 70 60 50 40 30 20 10 0 0

20

40

60

80

100

120

140

Time, (min.)

120

35

100

30 25

80

20 60 15 40

10

20

5

0

uranium uptake qe mg/g

uranium adsorpon, %

Figure 9. Equilibrium time versus the uranium adsorption efficiency (A%) on the RHA–alumina composite.

0 0

2

4

6

8

phase ratio Figure 10. The effect of the solid/liquid ratio on the uranium adsorption efficiency, A%, and uptake (mg/g) by the RHA–alumina composite.

0.1 g of the adsorbent was contacted with 10 ml of a 120.6 mg/l uranium solution with different concentrations of the desired elements at 25  2 C for 1 h, and the solution pH was maintained at 5. Figure 12 shows that in the presence of co-ions, the uranium uptake per cent on the RHA–alumina composite slightly decreased and exhibited the trend Thþ4> Feþ3> Cuþ2> Caþ2. This trend may explain why the adsorption sites present on the RHA–alumina composite are better for thorium, iron and copper ions than calcium.

Youssef et al.

13

Urani um adsorption efficency, (% A)

97

96

95

94

93

92 20

25

30

35

40

45

50

55

60

65

Temperature, (°C)

Uranium adsorption efficiency, (A%)

Figure 11. The effect of the temperature on the uranium adsorption efficiency (A%) of the RHA–alumina composite. 101

Ca+2

100

Cu+2

99 98 97

Fe+3

96 95 94

Th+4

93 0

20

40

60

80

100

120

140

Co-ions concentration (mg/L)

Figure 12. The effect of the presence of co-ions on the uranium adsorption efficiency, A% by RH–alumina composite.

Elution or desorption of uranium The expediency of using the AC doped with aluminium oxide for the preconcentration– separation of uranium was assessed by elution studies. For these studies, an amount of the adsorbent was shaken under the conditions previously determined to load the adsorbent. The desorption of the uranyl ions into the RHA–alumina composite was performed in a

14

Adsorption Science & Technology 0(0) 45

Elution efficiency, (% )

40 35 30 25 20 15 10 5 0

HNO3

HCl

H2SO4

NaCl

Na2CO3

Eluent type

Figure 13. The effect of the eluent agent on the desorption of uranium from the loaded RHA–alumina composite.

batch mode. The factors examined included the eluent agent, stripping agent concentration, phase ratio and shaking time.

The effect of the eluent type The following salt and acid solutions, Na2CO3, NaCl, H2SO4, HCl and HNO3, were used to desorb uranium from the loaded adsorbent. The elution experiments were conducted by shaking 0.05 g of the loaded adsorbent for 15 min with 10 ml of the elution agents (0.1 mol/l). The results are shown in Figure 13. HNO3 was the optimum solution for the elution of uranium from the loaded adsorbent. The other factors controlling the elution process that were studied were the shaking time, phase ratio and concentration of the elution agent.

The effect of the shaking time To study the influence of the equilibrium time on the uranium elution from the loaded adsorbent, a series of experiments were performed by contacting 0.05 g of the loaded adsorbent with 10 ml of 0.1 mol/l HNO3 for interval periods ranging from 15 to 120 min. The results are shown in Figure 14. The stripping efficiency increases with the increasing time from 15 to 60 min, but it did not further increase as the shaking time increased above 60 min. Therefore, 60 min is the optimum shaking time.

The effect of the phase ratio (S/a) The influence of the solid–liquid ratio on the desorption of U(VI) from the loaded RHA– alumina composite was varied (0.05 g/5 ml, 0.05 g/7.5 ml, 0.05 g/10 ml and 0.025 g/10 ml) at 298 K with 1 h shaking time and 0.1 mol/l HNO3. The results are illustrated in Figure 15, and they revealed that the uranium desorption or elution efficiency decreased due to dilution.

Youssef et al.

15

60

Elution efficiency,(%)

50

40

30

20

10

0 0

15

30

45

60

75

90

105

120

135

Shaking time, (min.)

Figure 14. The effect of the equilibrium time on the desorption of uranium from the loaded RHA–alumina composite. 120

Elution efficiency, (%)

100

80

60

40

20

0 0.05/5

0.05/7.5

0.05/10

0.025/10

Phase ratio, (S/A), (g/mL)

Figure 15. The effect of the phase ratio on the desorption of uranium from the loaded RHA–alumina composite.

The effect of the elution agent (HNO3) concentration To study the effect of the elution agent concentration (HNO3) on the uranium elution efficiency, various concentrations of HNO3 were prepared. Then, 10 ml of the HNO3 and 0.05 g of the loaded adsorbent were combined and shaken for 60 min. The results are shown

16

Adsorption Science & Technology 0(0) 120

Elution efficiency, (% )

100 80 60 40 20 0 0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1

HNO3 concentration, (M)

Figure 16. The effect of nitric acid concentration on desorption of uranium from loaded RHA–alumina composite.

in Figure 16. Obviously, 0.5 mol/l HNO3 can elute approximately 100% of the loaded uranium, and this was chosen as the optimum concentration.

Case study A case study was performed based on the sorption results using 500 ml of a waste solution with a uranium concentration of 100 mg/l. The solution was contacted with 5 g of the RHA– alumina composite for 60 min at room temperature 25  2 C and a pH of 5. After equilibration, the solution (case study) was filtered and analysed to determine the uranium concentration. The overall adsorption capacity of the RHA–alumina composite was 99%. The uranium desorbed with 0.5 M HNO3, and the uranium was precipitated by H2O2 after adjusting the pH to 2.5–3.

The experimental capacity of RHA compared with the desorption capacities of some chelate-modified solid-phase extraction procedures Table 4 shows different chemically modified adsorbents that are used for uranium adsorption. The sorption capacities (mg/g) were in the following order: RHA–alumina composite > AC> Coir pith> palm shell> orange peels> powdered corncob> sunflower> succinic acid-impregnated Amberlite XAD-4> date pits> natural clay> natural clinoptilolite zeolite (85 > 28.5 > 28 > 25 > 15.91 > 14.21 > 13.45 > 10 > 3.53 > 0.7 mg/g, respectively). When preparing an adsorbent of aluminium carbonate with silica and caustic soda (AS composite) in the same conditions as RHA–alumina composite prepared and applying optimum conditions for adsorption of uranium, the uptake was found to be 35 mg/g.

Youssef et al.

17

Table 4. Desorption capacities of some chelate-modified solid-phase extraction procedures for U(VI) uptake from aqueous solutions. Adsorbent RHA–aluminium composite Gel-amide AC Coir pith Palm shell Orange peels Powdered corncob Sunflower Succinic acid impregnated amberlite XAD-4 Date pits Natural clay in order Natural clinoptilolite zeolite

Adsorption capacity (mg/g) pH

Shaking Adsorption time % References

85 28.98 28.5 28 200 15.91 14.21 13.45 12.33

5 6 – 4.3 1 4 5 4 4.5

1h 6h – 0.5 h 3h 1 1

10 3.53 0.7

6–7 0.25 6 2 6 0.75

0.75

96.35 30% – – – – – – 80–90%

Present study Venkatesan et al. (2004) Abbasi and Streat (1994) Parab et al. (2005) Kushwaha and Sudhakar (2013) Mahmoud (2013) Mahmoud (2016) Roongtanakiat et al. (2010) Ahmad et al. (2015)

98

Saad et al. (2008) Eba et al. (2013) Camacho et al. (2010)

95.6

Conclusions RHA–alumina composite having a high surface area and large pore volume was synthesized. The structure of the new sorbent was characterized by FTIR spectroscopy and SEM. The RHA–alumina composite, which was synthesized by the ignition of RH with alumina and caustic soda at 500 C to have activation sites on the adsorbent surface, was used for the removal of U(VI) ions from a synthetic solution and a case sample study using a batch system, and it has been established as an effective adsorbent. The binding preference for uranium ions is attributed to its atomic properties and the chemistry of the solution, e.g. the pH. The proper pH value for the adsorption of the studied ions onto the RHA–alumina composite from an aqueous solution was 5. A higher adsorption capacity is obtained at ambient temperature 25  2 C. A Langmuir isotherm fits the equilibrium data better than the Freundlich isotherm, and the uranium adsorption capacity was 68 mg/l. In addition, the adsorption equilibrium data fit very well to both models in the studied concentration range. The results of the study indicate that the investigated RHA–alumina composite can effectively remove uranium anionic species from aqueous solutions. Declaration of Conflicting Interests The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Funding The author(s) received no financial support for the research, authorship, and/or publication of this article.

ORCID iD M. S. Hagag

http://orcid.org/0000-0002-3591-836X

18

Adsorption Science & Technology 0(0)

References Abbasi WA and Streat M (1994) Adsorption of uranium from aqueous solutions using activated carbon. Separation Science and Technology 27: 1217–1230. Adamska AM, Bright EL, Sutcliffe J, et al. (2015) Characterisation of electrodeposited polycrystalline uranium dioxide thin films on nickel foil for industrial applications. Thin Solid Films 597: 57–64. Afsari M, Safdari J, Towfighi J, et al. (2012) The adsorption characteristics of uranium hexafluoride onto activated carbon in vacuum conditions. Annals of Nuclear Energy 46: 144–151. Ahmad A, Siddique JA, Laskar AM, et al. (2015) New generation Amberlite XAD resin for the removal of metal ions: A review. Journal of Environmental Science 31: 104–123. Ahn S, Irukuvarghula S and McDeavitt SM (2016) Microstructure of a-U and d-UZr2 phase uranium– zirconium alloys irradiated with 140-keV Heþ ion-beam. Journal of Alloys and Compounds 681: 6–11. Amrute AP, Krumeich F, Mondelli C, et al. (2013) Depleted uranium catalysts for chlorine production. Chemical Science 4: 2209–2217. Bacci C, Boniface J, Bonino R, et al. (1989) Nuclear instruments and methods in physics research Section A: Accelerators, spectrometers. Detectors and Associated Equipment 279: 169–179. Bhatnagar A and Jain AK (2005) A comparative adsorption study with different industrial wastes as adsorbents for the removal of cationic dyes from water. Journal of Colloid and Interface Science 28: 49–55. Biswas SP, Pathak N and Roy SB (2012) Carrier facilitated transport of uranium across supported liquid membrane using dinonyl phenyl phosphoric acid and its mixture with neutral donors. Desalination 290: 74–82 Camacho LM, Deng S and Parra RR (2010) Uranium removal from groundwater by natural clinoptilolite zeolite: Effects of pH and initial feed concentration. Journal of Hazardous Materials 175: 393–398. Chegrouche S, Mellah A and Telmoune S (1997) Removal of lanthanum from aqueous solutions by natural bentonite. Water Research 31: 1733–1737. Chen Y, Wei Y, He L, et al. (2016) Separation of thorium and uranium in nitric acid solution using silica based anion exchange resin. Journal of Chromatography A 1466: 37–41. Dabrowski A (2001) Adsorption – From theory to practice. Advances in Colloid and Interface Science 93: 135–224. Dartiguelongue A, Chagnes A, Provost E, et al. (2016) Modelling of uranium(VI) extraction by D2EHPA/TOPO from phosphoric acid within a wide range of concentrations. Hydrometallurgy 165(1): 57–63. Daud WMAW, Ali WSW and Sulaiman MZ (2000) The effects of carbonization temperature on pore development in palm-shell based activated carbon. Carbon 38: 1925–1932. Dong Y, Liao W and Suo Z (2015) Uranium oxide-supported gold catalyst for water–gas shift reaction. Fuel Processing Technology 137: 164–169. Donia AMM, Atia AA, Ewais MMM, et al. (2009) Removal of uranium(VI) from aqueous solutions using glycidyl methacrylate chelating resins. Hydrometallurgy 95: 183–189. Dubinin MM (1960) The potential theory of adsorption of gases and vapors for adsorbents with energetically non-uniform surface. Chemical Reviews 60: 235–266. Eba F, Nlo JN, Ondo JA, et al. (2013) Batch experiments on the removal of U(VI) ions in aqueous solutions by adsorption onto a natural clay surface. Journal of Environment and Earth Science 3: 11–24. El Sayed MS (2003) Uranium extraction from gattar sulfate leach liquor using aliquat-336 in a liquid emulsion membrane process. Hydrometallurgy 68(1–3): 51–56. Fan F, Ding H, Bai J, et al. (2011) Sorption of uranium(VI) from aqueous solution onto magnesium silicate hollow spheres. Journal of Radioanalytical and Nuclear Chemistry 289: 367–375.

Youssef et al.

19

Franco DSP, Cunha JM, Dortzbacher GF, et al. (2017) Adsorption of Co(II) from aqueous solutions onto rice husk modified by ultrasound assisted and supercritical technologies. Process Safety and Environmental Protection 109: 55–62. Gajowiak A, Gładysz-Płaska A, Sternik D, et al. (2013) Sorption of uranyl ions on organosepiolite. Chemical Engineering Journal 219: 459. Ghoshal K, Kaity S, Mishra S, et al. (2014) Microstructural investigation of uranium rich U–Zr–Nb ternary alloy system. Geochimica et Cosmochimica Acta 446: 217–233. Grabias E, Gładysz-Płaska A, Ksia˛_zek A, et al. (2013) Efficient uranium immobilization on red clay with phosphates. Environmental Chemistry Letters 12: 297–301. Guo M, Yin X and Huang J (2017) Preparation of novel carbonaceous solid acids from rice husk and phenol. Materials Letters 196: 23–25. Hameed BH, Din ATM and Ahmad AL (2007) Adsorption of methylene blue onto bamboo-based activated carbon: Kinetics and equilibrium studies. Journal of Hazardous Materials 141: 819–825. Han R, Zou W, Wang Y, et al. (2007) Removal of uranium (VI) from aqueous solutions by manganese oxide coated zeolite: Discussion of adsorption isotherms and pH effect. Journal of Environmental Radioactivity 93: 127–143. Ho YS and McKay G (1999) The sorption of lead(II) ions on peat. Water Research 33: 578–584. Hore-Lacy I (2016) Uranium for nuclear power: Resources, mining and transformation to fuel. London, United Kingdom: Woodhead Publishing, pp.215–236. Hu Z and Srinivasan MP (1999) Preparation of high-surface-area activated carbons from coconut shell. Microporous and Mesoporous Materials 27: 11–18. Ilaiyaraja P, Deb AKS, Ponraju D, et al. (2017) Surface engineering of PAMAM-SDB chelating resin with Diglycolamic Acid (DGA) functional group for efficient sorption of U(VI) and Th(IV) from aqueous medium. Journal of Hazardous Materials 328: 1–11.  ˛tkowski A, Choma J, et al. (1991) Active Carbon. New York: Ellis Horwood. Jankowska H, Swia Kolev SD, St John AM and Cattrall RW (2013) Mathematical modeling of the extraction of uranium (VI) into a polymer inclusion membrane composed of PVC and di-(2-ethylhexyl) phosphoric acid. Journal of Membrane Science 425–426: 169–175. Kushwaha S and Sudhakar P (2013) Sorption of uranium from aqueous solutions using palm-shellbased adsorbents: A kinetic and equilibrium study. Journal of Environmental Radioactivity 116: 115–124. Lee JH, Kwon JH, Lee J, et al. (2017) Preparation of high purity silica originated from rice husks by chemically removing metallic impurities. Journal of Industrial and Engineering Chemistry 50: 79–85. Mahmoud AM (2013) Removal of uranium (VI) from aqueous solution using low cost and eco-friendly adsorbents. Journal of Chemical Engineering & Process Technology 4: 6–10. Mahmoud AM (2016) Kinetics studies of uranium sorption by powdered corn cob in batch and fixed bed system. Journal of Advanced Research 7: 79–89. Marczenko Z (1976) Spectrophotometric Determination of Elements. New York: Ellis Horwood Ltd, Halsted Press, p.580. Mellah A and Chegrouche S (1997) The removal of zinc from aqueous solutions by natural bentonite. Water Research 31: 621–629. Mohamed RM, Khalid IA and Barakat MA (2015) Rice husk ash as a renewable source for the production of zeolite NaY and its characterization. Arabian Journal of Chemistry 8: 48–53. Ogden MD, Moon EM, Wilson A, et al. (2017) Application of chelating weak base resin Dowex M4195 to the recovery of uranium from mixed sulfate/chloride media. Chemical Engineering Journal 317: 80–89. Parab H, Joshi S, Shenoy N, et al. (2005) Uranium removal from aqueous solution by coir pith: Equilibrium and kinetic studies. Bioresource Technology 96: 1241–1248. Roongtanakiat N, Sudsawad P and Ngernviji N (2010) Uranium absorption ability of sunflower, vetiver and purple guinea grass. Kasetsart Journal – Natural Science 44: 182–190.

20

Adsorption Science & Technology 0(0)

Saad EM, Mansour RA and El-Shahawi MS (2008) Sorption profile and chromatographic separation of uranium (VI) ions from aqueous solutions onto date pits solid sorbent. Talanta 76(5): 1041–1046. Setthaya N, Chindaprasirt P, Yin S, et al. (2017) TiO2-zeolite photocatalysts made of metakaolin and rice husk ash for removal of methylene blue dye. Powder Technology 313: 417–426. Shi L, Zhu P, Yang R, et al. (2017) Functional rice husk as reductant and support to prepare as-burnt Cu-ZnO based catalysts applied in low-temperature methanol synthesis. Catalysis Communications 89: 1–3. Sinyoung S, Kunchariyakun K, Asavapisit S, et al. (2017) Synthesis of belite cement from nano-silica extracted from two rice husk ashes. Journal of Environmental Management 190: 53–60. Sobrosa FZ, Stochero NP, Marangon E, et al. (2017) Development of refractory ceramics from residual silica derived from rice husk ash. Ceramics International 43: 7142–7146. St John AM, Cattrall RW and Kolev SD (2012) Transport and separation of uranium(VI) by a polymer inclusion membrane based on di-(2-ethylhexyl) phosphoric acid. Journal of Membrane Science 409–410: 242–250. Venkatesan KA, Sukumaran VM, Antony P, et al. (2004) Extraction of uranium by amine, amide and benzamide grafted covalently on silica gel. Journal of Radioanalytical and Nuclear Chemistry 260: 443–450. Yakout SM, Metwally SS and El- Zakla T (2013) Uranium sorption onto activated carbon prepared from rice straw: Competition with humic acids. Applied Surface Science 280: 745–750. Zhu Z, Pranolo Y and Cheng CY (2016) Uranium recovery from strong acidic solutions by solvent extraction with Cyanex 923 and a modifier. Minerals Engineering 89: 77–83.