Adsorption of Cefocelis Hydrochloride on Macroporous Resin: Kinetics

May 4, 2016 - Japanese Pharmacopoeia Fourteenth; The Ministry of Health Labour and Welfare: Tokyo, 2001, 334−335. (23) Weber, T. W.; Chakravorti, R...
0 downloads 0 Views 770KB Size
Article pubs.acs.org/jced

Adsorption of Cefocelis Hydrochloride on Macroporous Resin: Kinetics, Equilibrium, and Thermodynamic Studies Feng Xue, Yingting Xu, Shizhao Lu, Shengui Ju,* and Weihong Xing College of Chemical Engineering, Nanjing Tech University, Nanjing, 210009, People’s Republic of China ABSTRACT: The adsorptions of Cefocelis hydrochloride in aqueous solution on macroporous resin (HP-20) were studied. On the basis of static experiments with HP-20 resin, the adsorption kinetics, isotherms, and thermodynamics of Cefocelis hydrochloride in aqueous solution on macroporous resin (HP-20) were investigated. Adsorption equilibrium data were correlated with the Langmuir and Freundlich equations. Adsorption data of kinetic were modeled using the Crank equation and fitting for the diffusion coefficient (De), and the value of the liquid film mass transfer coefficient (kf) was estimated from Carberry equation. Dynamic adsorption was performed on HP-20 resin packed in a glass column to obtain optimal parameters. The initial concentration, bed height, and residence time were considered to determine the dynamic test for the breakthrough curve. The values of thermodynamic parameters including the isosteric adsorption enthalpy (ΔH), free energy (ΔG), entropy (ΔS), and adsorption activation energy (Ea) demonstrated the process of adsorption was spontaneous and exothermic. The adsorption process was controlled by a physical mechanism.



sion), and the internal-surface diffusion of the adsorbent.9,10 Casillas et al.11 studied the kinetics model using the linear adsorption isotherm and a single effective internal diffusion coefficient. They fitted the effective Sherwood number employed by the agitation method against the experiments. Yang et al.12 identified the significance of pore and surface diffusions and estimated the values by the nonlinear regression and “the half-time” methods. However, the diffusion coefficient needs to be further simplified in engineering application. The rate of the internal-surface diffusion is normally very fast and has less influence on the adsorption process. To simplify the modeling in the adsorption process, the rate of the external diffusion is eliminated by shanking in the adsorption process of the experiment,11 therefore the internal-particle diffusion is considered as the kinetically controlled step when compared with external diffusion for the adsorption process. The internalparticle diffusion is estimated by fitting experiment data by the Crank model.13 Furthermore, the single effective diffusion coefficient in intraparticle diffusion is used for the description of the mass transfer in the adsorbent.14,15 There are several methods that are used to estimate the film transfer coefficient (kf).14−16 For 0.1 < Re/ε < 1000, the Carberry equation16 is utilized to calculate the liquid film mass transfer coefficient (kf). Adsorption isotherms are important for the description of the interaction between adsorbed molecules or ions and adsorbent surface sites. The Langmuir and Freundlich equations are the popular models for the simulation of

INTRODUCTION Cefoselis sulfate is a novel, parenteral, and fourth-generation cephalosporin. It has a broad spectrum of antibacterial activity against Gram-positive and Gramnegative bacteria, including Pseudomonas aeruginosa.1−3 Cefoselis hydrochloride (Figure 1)

Figure 1. Chemical structure of Cefoselis hydrochloride.

is the key intermediate to prepare Cefoselis sulfate.4−6 The purity, efficacy, and safety performance of Cefoselis sulfate are directly influenced by the purity of the intermediate.7 Adsorption by macroporous resin has the advantages including high absorption of targeted components, low operating temperature, (which ensures the stability of thermosensitive cephalosporin compounds) and feasibility to desorption for further separation and purification. Adsorption kinetics can describe the adsorption process of Cefocelis hydrochloride quantitatively. It is crucial for designing adsorption facilities and understanding the adsorption types and mechanisms.8 There are three types of diffusions of the molecules in the macroporous resin, including external diffusion (film diffusion), intraparticle diffusion (pore diffu© XXXX American Chemical Society

Received: February 26, 2016 Accepted: April 26, 2016

A

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Some Properties of HP-20 Resin resin

polymer matrix

polarity

bead size/(mm)

pore diameter/(nm)

BET surface area/(m2/g)

specific gravity/(g/mL)

HP-20

styrene

nonpolar

0.783

93

500−600

1.03−1.07

flowed through the column in 15 min with the flow rate of 1.0 mL/min. Photodiode array detection was used to detect Cefocelis hydrochloride at the wavelength of 254 nm.

adsorption isotherms in low concentration regions. The former equation describes the monolayer adsorption on the surface, and the latter equation describes the monolayer adsorption as well as multilayer adsorption.17 For adsorption thermodynamics, the thermodynamic parameters including free energy change, enthalpy change, entropy change, and active energy of adsorption can provide in-depth information on inherent energetic changes that are associated with adsorption.18,19 Hasan et al.19 calculated the active energy from the Arrhenius equation using the pseudo-second-order constant. The kinetics, isotherms, and thermodynamics on the adsorption of Cefocelis hydrochloride have not been investigated up to now. In this current study, we studied the isotherm model by the Langmuir and Freundlich models to simulate the adsorption process. On the basis of the kinetics experiment data of adsorption of Cefocelis hydrochloride on HP-20 resin, the diffusion coefficient (De) was fitted by the Crank model,13 and the liquid film mass transfer coefficient (kf) was estimated by the Carberry equation.16 Column dynamic performance of absorption Cefocelis hydrochloride on HP-20 resin was investigated by breakthrough curve under experimental conditions. Basic thermodynamic parameters, thermodynamic function relationships, and adsorption isotherm equations were obtained. Our research can lay the foundation for a theoretical basis of Cefocelis chloride on adsorption resin for industry application.

2. EQUATION AND CLASSICAL MODELS 2.1. Equation and Dynamic Models. The following equation was used to quantify the capacity of adsorption. qe = V (Co − Ce)/w E(%) =

(1)

Co − Ce 100 Co

(2)

where qe is the adsorption quantity at equilibrium (mg/g); Co and Ce are the initial and equilibrium concentrations (mg/mL), respectively; w is the weight of the wet resin (g). The following equation was used to quantify the ratio of desorption (D): D(%) =

Cd × Vd 100 (Co − Ce)Vi

(3)

where Cd is the concentration of the solute in the desorption solution (mg/mL), Vi is the volume of initial sample solution (mL), Vd is the volume of the eluent (mL). Langmuir equation: qe =

1. MATERIALS AND METHODS 1.1. Reagents, Materials, and Apparatus. Cefoselis hydrochloride was prepared in our laboratory with a purity of 97.5%.4−6 HPLC-grade methanol for preparation of mobile phase was purchased from Sigma-Aldrich. Other reagents including disodium hydrogen phosphate dodecahydrate, potassium dihydrogen phosphate, and methanol are of analytical grade and were purchased from Sino pharm Chemical Reagent Co., Ltd. Deionized water was purchased from Wahaha Company. Dianon HP-20 resin was purchased from Mitsubishi Chemical Co., Ltd. Some of its properties are summarized in Table 1. We compared the HP-20 resin with the other resins XR-935C and XR-925C, which were purchased from Shanghai Hua Zhen Science and Technology Co., Ltd. HP-20 resin displayed higher adsorption capacity, adsorption, and desorption ratios than XR-935C and XR-925C. Therefore, we selected HP-20 resin as adsorbent for the purification of Cefocelis hydrochloride. The resin was pretreated with 1 M HCl and NaOH solutions successively to remove monomers and porogenic agents trapped inside the pores during the synthesis process.20 Then the resin was soaked in methanol for at least 24 h. Prior to the adsorption experiments, the methanol in a certain amount of resin was thoroughly replaced with deionized water.21 1.2. Analytical Procedures. The HPLC method22 on a Wufeng liquid chromatographic system equipped with an ODS3 C18 column (GL Science Co. Ltd., 180 × 4.6 mm i.d., 5.0 μm) was used to determine the composition of samples. The column oven temperature was set at 295 K. The mobile phase consisted of methanol and deionized water (20:80, v/v) and

qm × b × Ce (1 + b × Ce)

(4)

Freundlich equation: qe = K × Ce1/ n

(5)

where b is the Langmuir constant; qm is the theoretically calculated maximum adsorption capacity (mg/g resin); K is the Freundlich constant; n is an empirical constant. The Crank’s single-hole diffusion model was used to fit the data of static dynamic experiment data.13 E=

⎡6 Mt =1−⎢ 2 M∞ ⎣⎢ π



∑ n=1

⎛ n2π 2Dt ⎞⎤ 1 e ⎥ ⎜− ⎟ exp n2 r 2 ⎠⎥⎦ ⎝

(6)

where Mt represents the adsorption capacity of t at the moment (mg/g); M∞ represents the adsorption capacity at the end of time (mg/g); De represents the diffusion coefficient (cm2/s); r represents the particle size of the adsorbent (cm). Equation 6 can converge rapidly as t is larger because the sum of the high order terms of the series is very small. Therefore, the error is less than 2% when (De × t/r2) < 0.1, taking that the first approximation is sufficient. When n = 1, eq 6 was simplified to M t /M∞ = 1 −

⎛ π 2D t ⎞ 6 exp⎜ − 2 e ⎟ 2 r ⎠ π ⎝

(7)

The film transfer coefficient might be calculated from the following correlation, developed by Carberry:16 (ε × k f /μ) × Sc 2/3 = 1.15 × (Re/ε)−0.5 B

(8)

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

breakthrough volume. Breakthrough volumes were calculated at 1% of the maximum absorbance of the adsorbates.28 The concentration of Cefocelis hydrochloride was analyzed by spectrophotometry. 3.2. Thermal Stability of Cefocelis Hydrochloride Tests. Cefocelis hydrochloride solution was placed at 295 and 308 K for different times, and the assays of Cefocelis hydrochloride at different treated times were determined by spectrophotometry to monitor their degradation rates. 3.3. Static Adsorption Tests. Cefocelis hydrochloride adsorption data on HP-20 resin were determined following the literature:29 0.2 g of hydrated resin was added into each flask with 50 mL of Cefocelis hydrochloride aqueous solution with different initial concentrations (0.04, 0.08, 0.12, 0.16, and 0.2 mg/mL). The flasks were shaken on a shaker with 100 run per minute (rpm) at different temperatures for 12 h. 3.4. Adsorption Kinetics Tests. The adsorption kinetics of the HP-20 macroporous resin was studied by contacting 100 mL of Cefocelis hydrochloride with 5 g of hydrated resin in a shaker solution. The compositions of the liquid at different times were determined by HPLC when adsorption equilibration was achieved. 3.5. Column dynamic adsorption tests. Cefocelis hydrochloride solution was pumped through each glass column packed with wet HP-20 resin at 295 K, and then was executed at three following conditions: (i) under different initial concentrations of 0.2 and 0.1 mg/mL with the same bed height of 20.5 cm and the same injection speed of 35 mL/h; (ii) under different bed heights of 16.5 and 20.5 cm with the same initial concentration of 0.2 mg/mL and the same injection speed of 35 mL/h; (iii) under different residence time of 10 and 17.5 min with the same initial concentration of 0.2 mg/mL and the same bed height of 20.5 cm.

In which the Reynolds and Schmidt numbers were calculated from eqs 9 and 10. Re = d p × u × ρ /μ

(9)

μ ρ × De

(10)

Sc =

where μ represents the fluid viscosity (g/(cm s); ρ represents fluid density (g/cm3); De represents the diffusion coefficient (cm2/s). Equation 8 was rewritten as, u k f = × Sc(−2/3) × 1.15 × Re−0.5 (11) ε where u represents empty column flow rate (m/s); ε represents bed porosity, and Re represents Reynolds number. 2.2. Adsorption Thermodynamic Properties. Isosteric adsorption enthalpy (ΔH) could be calculated by the Van’t Hoff equation.23 ln Ce = −ln Ko + ΔH /(RT )

(12)

where R is the gas contant of 8.314 (J/(mol K)); T is the thermodynamic temperature (K); Ko is a constant. Adsorption free energy (ΔG) could be calculated by Gibbs equation.24 ΔG = −RT

∫0

x

(qe /X ) dX

(13)

where X is the molar fraction of adsorbate in equilibrium in solution. Equation 13 is simplified to the following eq 14 when the adsorption process follows the Freundlich model in eq 4:

ΔG = −nRT

(14)

Entropy of adsorption (ΔS) could be calculated by Gibbs− Helmholtz equation.25 ΔS = (ΔH − ΔG)/T

4. RESULTS AND DISCUSSION 4.1. Adsorption and Desorption Properties of HP-20 Resin. The adsorption and desorption performances were associated with chemical features of the solution and the adsorbents.30 Resin with similar polarity to solvent exhibited good desorption abilities. The adsorption capacity of this HP20 resin for Cefocelis hydrochloride, adsorption, and desorption ratios were tested to be 4.55 mg/g, 75.4%, and 78.2%, respectively. These results indicated HP-20 resin possessed a strong affinity for Cefocelis hydrochloride, and the process of adsorption was reversible. Consequently, HP-20 resin could be used as adsorbent in the adsorption process of separating Cefocelis hydrochloride. In addition, taking the nonpolar properties of HP-20 resin and polarity of the Cefocelis hydrochloride into consideration, the desorption results suggested that Cefocelis hydrochloride was probably the polar compound based on the theory of similarity and intermiscibility.27 4.2. Thermostability of Cefocelis Hydrochloride. Cephalosporin compounds are known as heat-sensitive compounds. The chemical structure of Cefoselis hydrochloride is not stable at high temperatures. Therefore, it is necessary to modify the operation temperature in the purification process. Figure 2 shows the dependence of the purities of Cefocelis hydrochloride on time at 295 and 308 K, respectively. The purity of Cefocelis hydrochloride decreased by only 0.7% at 295 K after 18 h exposure. The decomposition of Cefocelis hydrochloride could be ignored in our operation process because our operation time was less than 12 h. The

(15)

Adsorption activation energy (Ea) was calculated by Arrhenius equation.26 ⎛ E ⎞ De = Do exp⎜ − a ⎟ ⎝ RT ⎠

(16)

where De is the diffusion coefficient; Do is proportional coefficient and is related to the properties of the adsorbent and adsorbate and the temperature. De is fitted to eq 7 according to the Crank model.13

3. EXPERIMENTAL PROCEDURES 3.1. Adsorption and Desorption Properties of HP-20 Resin. To acquire the adsorption and desorption properties, static adsorption and desorption tests were performed. In the adsorption experiment,27 2 g of resin (wet weight) and 50 mL of methanol (initial concentration of Cefocelis hydrochloride was 0.2 mg/mL) were added into flask. The flask was shaken (120 rpm) in an oscillator at 295 K for 12 h, and then the concentration of Cefoselis hydrochloride in the liquid phase was determined by spectrophotometry. The desorption study28 was performed as follows: After reaching adsorption equilibrium, the HP-20 resin particles were separated from the solution by filtration, and washed with deionized water. The cleaned resin particles were packed in a glass column. An aqueous solution of isopropyl alcohol (10:90, v/v) was pumped through the column to measure the C

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 2. Time-assay curves of Cefocelis hydrochloride at 295 and 308 K.

Figure 4. Langmuir isotherms of adsorption Cefocelis hydrochloride on HP-20 resin at different temperatures, at a concentration of 0.2 mg/mL.

decomposition rate increased to 9.3% after 18 h while the operation temperature increased to 308 K. The decomposition resulted from the ring opening reaction of the beta-internal amide ring in Cefocelis hydrochloride. These results indicated that operation temperature had a great influence on the stability of Cefocelis hydrochloride. The purification experiments were carried out at room temperature, 295 K, considering the operation convenience. 4.3. Static Adsorption Curve of Cefocelis Hydrochloride on HP-20 Resin. The adsorption constants together with the correlation coefficients were obtained using Langmuir and Freundlich curve fittings in Origin 8.6. The parameter values of the two models are listed in Table 2. Figures 3 and 4

Table 3. Characteristic Parameters of the Adsorbent Bed Layer and Operation Data Bed Parameters and Calculation Data bed height L/m bed inner diameter Dz/m particle diameter dp/mm packing density kg/m3 bed porosity ε particle porosity εp empty column flow rate u (m/s) Reynolds number Re Schmidt number Sc (105)

Table 2. Parameters of Langmuir and Freundlich Models Freundlich model

HP-20 resin at 275 and 295 K as shown in Figure 5. The model fit well with the experimental data. De values were calculated by

Langmuir model

T/K

K

n

R2

qm

b

R2

275 285 295

132.56 109.77 96.03

1.270 1.265 1.250

0.9977 0.9900 0.9993

26.02 25.38 25.56

15.59 12.50 9.88

0.9977 0.9984 0.9993

0.205 0.006 0.315−1.25 (average value 0.783) 1030−1070 0.48 0.562 0.02 174 2.31 at 275 K and 1.07 at 295 K

Figure 5. Fitting results of adsorption data by Crank model with a Cefocelis hydrochloride concentration of 0.2 mg/mL. Figure 3. Freundlich isotherms of Cefocelis hydrochloride on HP-20 resin at different temperatures, at a concentration of 0.2 mg/mL.

eq 6 as 3.90 × 10−9 and 8.39 × 10−9 cm2/s at 275 and 295 K, respectively. The correlation coefficients (R2) were obtained by eq 7 as 0.8383 and 0.8617 at 275 and 295 K, respectively. 4.4.2. Liquid Film Mass Transfer Coefficient (kf). The kf values estimated by the flow velocity of the empty bed could simplify the calculation of the data and provide useful data for the design of the fixed bed. As eq 11 described, kf is related to the bed parameters and Reynolds and Schmidt numbers. Table 3 shows the bed and operation parameters for the empty bed. Reynolds and Schmidt numbers were accordingly calculated by eqs 9 and 10. Therefore, the kf value was estimated by eq 11 as 8.25 × 10−5 and 1.38 × 10−4 cm/s at 275 and 295 K, respectively.

compare the two models with the adsorption experimental dependence of the concentration at different temperatures. The two models predicted the adsorption amounts under the investigated test conditions well. The variances (R2) of the modeling for the experiments under different temperatures were higher than 0.99 using the two models as shown in Table 3. These results further indicated that the adsorption process was mostly a monomolecular layer adsorption. 4.4. Adsorption Kinetics Test. 4.4.1. Diffusion Coefficient (De). The Crank model [eqs 6 and 7] was used to predict the adsorption rates of Cefocelis hydrochloride on the D

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

4.5. Column Kinetics of Adsorption of the Adsorbed Resin. Breakthrough curve is an important characteristic for the operating process for the adsorption curve. It reflects the adsorption equilibrium relationship between mobile phase and fixed phase. Adsorption kinetics and mass transfer mechanism of the adsorption process are the main basis of equipment design and operation. The initial concentration of Cefocelis hydrochloride, the residence time of the Cefocelis hydrochloride solution in the bed, and the height of the adsorbent bed are investigated to have effects on the breakthrough curve. 4.5.1. Effect of Initial Concentration on the Breakthrough Curve. Figure 6 shows the time course of the Ci/Co ratio under

increase in bed height was accompanied by a larger pressure drop, and consequently, the suitable bed height was selected as 20.5 cm.17 4.5.3. Effect of Residence Time on Adsorption Kinetics. The effect of residence time on dynamic adsorption capacity is shown in Figure 8. The breakthrough rate and utilization rate of

Figure 8. Breakthrough curves of Cefocelis hydrochloride on HP-20 at different resident times: the initial concentration is 0.2 mg/mL, bed height is 20.5 cm, and adsorption temperature is 295 K.

the adsorbents decreased when the residence time reduced from 17.5 to 10 min. In general, shortening the residence time had a negative effect on dynamic adsorption capacity of adsorbate because adsorbate molecules had no sufficient time to undergo interactions with the active sites at the surface of resin and vice versa.17 4.6. Adsorption Thermodynamic Properties. 4.6.1. Isosteric Adsorption Enthalpy (ΔH), Free Energy (ΔG), and Entropy (ΔS). Figure 9 shows the plots of −ln Ce against 1/T

Figure 6. Breakthrough curves of Cefocelis hydrochloride on HP-20 at different initial concentration adsorption; bed height is 20.5 cm, injection speed is 35 mL/h, and adsorption temperature is 295 K.

the initial concentrations of 0.1 and 0.2 mg/mL on HP-20 resin. The adsorption rate of the adsorbent in the unit increased as the initial concentration increased, which resulted in the shortening of the breakthrough time. 4.5.2. Effect of Bed Height on Adsorption Kinetics. At certain initial concentrations of adsorbate, the breakthrough time was also affected by bed height. Figure 7 illuminates the

Figure 9. Determination of isosteric adsorption enthalpy of Cefocelis hydrochloride on HP-20.

according to Freundlich isotherm model in eq 12. The equilibrium concentrations Ce under different concentrations of 3, 5, and 7 mg/L were obtained using eq 1. The equivalent adsorption enthalpy was calculated from the slope of the line in Figure 9 and listed in Table 4. The Freundlich model fit linearly with the experimental data. R2 values for different experiments at different temperatures of 275, 285, and 295 K were as high as

Figure 7. Breakthrough curves of Cefocelis hydrochloride on HP-20 at different fixed-bed heights: concentration of Cefocelis hydrochloride is 0.2 mg/mL, injection speed is 35 mL/h, and adsorption temperature is 295 K.

time course of the Ci/Co ratio using bed heights of 16.5 and 20.5 cm, respectively. The penetration rate for a shorter bed (16.5 cm) was slightly faster than that of a longer bed (20.5 cm). The reason was that the improved bed layer increased the mass transfer area and enhanced the utilization of the adsorbent. Adsorption capacity showed a significant increase when the bed height was raised from 16.5 to 20.5 cm. A further

Table 4. Thermodynamic Properties of the Systems Tested

E

T/K

275 K

285 K

295 K

ΔH/(kJ/mol) ΔG/(kJ/mol) ΔS/[kJ/(mol K)]

−4.44 −1.33 0.013

−4.50 −3.00 0.0053

−4.60 −3.01 0.054

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Innovation Project of Jiangsu Province (CXZZ13_0451) for the provision of financial support.

0.988, 0.9981, and 0.9975, respectively, which suggested that eq 12 was suitable to be applied for the enthalpy change calculation. The negative value of ΔH implied the exothermic nature of the adsorption process,31 which suggested that qe decreased when T increased.27 In general, exothermic heat was required by an adsorption process to compensate for loses in entropy.32 Subsequently, with an increase in adsorption capacity, adsorption occurred on the sites of adsorbent with less activity. In addition, the absolute value of ΔG for each sample was lower than 40 kJ/mol, which indicated that the adsorption process was controlled by a physical mechanism.22 This result showed that the interaction between resin and Cefocelis hydrochloride was mainly electrostatic (Coulombic interactions).33 ΔG and ΔS could be calculated by eqs 14 and 15. The result was presented in Table 3. The negative value of ΔG confirmed the feasibility of the present adsorption process and the spontaneous nature of the adsorption.34 The positive values of ΔS at different temperatures reflected the good affinity of Cefocelis hydrochloride on HP-20 resin and the increased randomness at the solid/solution interface during the adsorption of Cefocelis hydrochloride onto HP-20 resin.35 4.6.2. Adsorption Activation Energy (Ea). Arrhenius eq 16 describes the relation of adsorption activation energy and the diffusion coefficient De. The diffusion coefficient De was already obtained and increased with the temperature increasing in section 4.4.1. However, the adsorption capabilities decreased with the increase of temperature. These results suggested that the adsorption process was exothermic. The activation energy Ea was calculated from the eq 16 as 25.83 kJ/mol. The magnitude of the activation energy (Ea) may give information about the type of adsorption process which can be physical or chemical. The physical adsorption process has lower activation energy values (5−40 kJ/mol) while the chemical adsorption has higher activation energy values (40−800 kJ/ mol).36 Our activation energy Ea was included in the reported range of physical adsorption,36 indicating that the adsorption of Cefocelis hydrochloride on HP-20 resin was a physicaladsorption controlled process.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are thankful to Professor R. F. Zhou for his suggestions and help in technical support during experiment work.



(1) Kuriyama, T.; Karasawa, T.; Nakagawa, K.; Nakamura, S.; Yamamoto, E. Antimicrobial susceptibility of major pathogens of orofacial odontogenic infection to 11 beta-lactam antibiotics. Oral Microbiol. Immunol. 2002, 17, 285−289. (2) Climo, M. W.; Markowitz, S. M.; Williams, D. S. Comparison of the in-vitro and in-vivo efficacy of FK037, vancomycin, imipenem and nafcillin against staphylococcal species. J. Antimicrob. Chemother. 1997, 40, 59−66. (3) Giamarellos-Bourboulis, E. J.; Grecka, P.; Tsitsika, A. In-vitro activity of FK037 (Cefoselis), a novel 4th generation cephalosporin, compared to cefepime and cefpirome on nosocomial staphylococci and gram-negative isolates. Diagn. Microbiol. Infect. Dis. 2000, 36, 185− 191. (4) Sakane, K.; Kawabata, K.; Miyai, K. New cephem compound and a process for preparation thereof. EP. Patent 0307804A2, 1998. (5) Ohki, H.; Kawabata, K.; Okuda, S.; Kamimura, T.; Sakane, K. FK037: a new parenteral cephalosporin with a broad antibacterial spectrum: synthesis and antibacterial activitu. J. Antibiot. 1993, 46, 359−361. (6) Sakane, K.; Kawabata, K.; Ohki, H. Process for the preparation of cephalosporin compound. JP Patent 92173792, 1992. (7) Zalewski, P.; Cielecka-Piontek, J.; Jelińska, A. Stability of Cefoselis sulfate in aqueous solutions. React. Kinet., Mech. Catal. 2013, 108, 285−292. (8) Ding, L.; Deng, H.; Wu, C.; Han, X. Affecting factors, equilibrium, kinetics and thermodynamics of bromide removal from aqueous solutions by MIEX resin. Chem. Eng. J. 2012, 181−182, 360− 370. (9) Sarici-Ozdemir, C.; Onal, Y. Equilibrium, kinetic and thermodynamic adsorptions of the environmental pollutant tannic acid onto activated carbon. Desalination 2010, 251, 146−152. (10) Ho, Y. S.; Mckay, G. Competitive sorption of copper and nickel ions from aqueous solution using peat. Adsorption 1999, 5, 409−417. (11) Casillas, J. L.; Martinez, M.; Addo-Yobo, F.; Aracil, J. Modelling of the adsorption of cephalosporin C on modified resins in a stirred tank. Chem. Eng. J. 1993, 52, 71−75. (12) Yang, S. A.; Pyle, D. L. The adsorption kinetics of cephalosporin C on non-ionic polymeric macropore Amberlite XAD-16 resin. J. Chem. Technol. Biotechnol. 1999, 74, 216−220. (13) Crank, J. The Mathematics of Diffusion; Oxford University Press: Oxford, 1956. (14) Ye, Z. H. Chemical Adsorption and Separation Process; China Petrochemical Press: Beijing, 1992. (15) Ruthven, D. M. Principles of Adsorption and Adsorption Processes; Wiley: New York, 1984. (16) Carberry, J. J. A boundary-layer model of fluid-particle mass transfer in fixed beds. AIChE J. 1960, 6, 460−463. (17) Ma, C. Y.; Tao, G. J.; Tang, J.; Lou, Z. X.; Wang, H. X.; Gu, X. H.; Hu, L.; Yin, M. L. Preparative separation and purification of rosavin in Rhodiola rosea by macroporous adsorption resins. Sep. Purif. Technol. 2009, 69, 22−28. (18) Liu, F.; Li, L.; Ling, P.; Jing, X.; Li, C.; Li, A.; You, X. Interaction mechanism of aqueous heavy metals onto a newly synthesized IDAchelating resin: isotherms, thermodynamics and kinetics. Chem. Eng. J. 2011, 173, 106−114.

5. CONCLUSIONS R2 of the Langmuir and Freundlich model were higher than 0.99, indicating that the adsorption of Cefocelis hydrochloride on HP-20 resin was a monolayer adsorption process. Then, the De values were obtained as 3.90 × 10−9 and 8.39 × 10−9 cm2/s at 275 and 295 K, respectively. The kf values were estimated as 8.25 × 10−5 and 1.38 × 10−4 cm/s at 275 and 295 K, respectively. Adsorption thermodynamic parameters including ΔH, ΔG, ΔS, and Ea were calculated, which all confirmed the nature of physical sorption process. Finally, the breakthrough curve of the dynamic test was determined by the different residence times, initial concentrations, and bed heights of adsorption of Cefocelis hydrochloride on HP-20 resin, providing some basic data for industrial design.



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*Tel: +86 25 83587182. E-mail: [email protected]. Funding

The authors wish to express sincere gratitude to the National Natural Science Foundation of China (21176118) and F

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(19) Hasan, M.; Ahmad, A. L.; Hameed, B. H. Adsorption of reactive dye onto crosslinked chitosan/oil palm ash composite beads. Chem. Eng. J. 2008, 136, 164−172. (20) Juang, R. S.; Shiau, J. Y. Adsorption isotherms of phenols from water onto macroreticular resins. J. Hazard. Mater. 1999, 70, 171−183. (21) Ma, C. Y.; Tang, J.; Wang, H. X.; Gu, X. H.; Tao, G. J. Simultaneous determination of six active compounds in Rhodiola L. by RP-LC. Chromatographia 2008, 67, 383−388. (22) National Institute of Health Sciences. Cefocelis Sulfate, The Japanese Pharmacopoeia Fourteenth; The Ministry of Health Labour and Welfare: Tokyo, 2001, 334−335. (23) Weber, T. W.; Chakravorti, R. K. Pore and solid diffusion models for fixed-bed adsorbers. AIChE J. 1974, 20, 228−238. (24) Sharma, U. Y. C.; Sinha, A. S. K.; Upadhyay, S. N.; Uma. Characterization and adsorption studies of Cocos nucifera L. Activated carbon for the removal of methylene blue from aqueous solutions. J. Chem. Eng. Data 2010, 55, 2662−2667. (25) John, P. B.; Marios, T. Removal of hazardous organic pollutants by biomass adsorption. J. Water. Pollut. Control Fed. 1987, 59, 191− 198. (26) Blondeau, P.; Tiffonnet, A. L.; Allard, E.; Haghighat, F. Physically based modelling of the material and gaseous contaminant interactions in buildings: models, experimental data and future developments. Adv. Build. Energy Res. 2008, 2, 57−93. (27) Chen, Y.; Zhang, D. Adsorption kinetics, isotherm and thermodynamics studies of flavones from vaccinium bracteatum thunb leaves on NKA-2 resin. Chem. Eng. J. 2014, 254, 579−585. (28) Jung, M. W.; Ahn, K. H.; Lee, Y.; Kim, K. P.; Paeng, I. R.; Rhee, J. S. Evaluation on the adsorption capabilities of new chemically modified polymeric adsorbents with protoporphyrin IX. J. Chromatogr. A 2001, 917, 87−93. (29) Hosseini-Bandegharaei, A.; Hosseini, M. S.; Sarw-Ghadi, M.; Zowghi, S.; Hosseini, E.; Hosseini-Bandegharaei, H. Kinetics, equilibrium and thermodynamic study of Cr(VI) sorption into toluidine blue o-impregnated XAD-7 resin beads and its application for the treatment of wastewaters containing Cr(VI). Chem. Eng. J. 2010, 160, 190−198. (30) Liu, Y.; Liu, J.; Chen, X.; Liu, Y.; Di, D. Preparative separation and purification of lycopene from tomato skins extracts by macroporous adsorption resins. Food Chem. 2010, 123, 1027−1034. (31) Cheng, Y. M.; Jin, X. H.; Gao, D.; Xia, H. F.; Chen, J. H. Thermodynamics and kinetics of lysozyme adsorption onto two kinds of weak cation exchangers. Biotechnol. Bioprocess Eng. 2013, 18, 950− 955. (32) Koyuncu, H.; Kul, A. R.; Yildiz, N.; Calimli, A.; Ceylan, H. Equilibrium and kinetic studies for the sorption of 3-methoxybenzaldehyde on activated kaolinites. J. Hazard. Mater. 2007, 141, 128−139. (33) Wang, X.; Deng, R.; Jin, X.; Huang, J. Gallic acid modified hyper-cross-linked resin and its adsorption equilibria and kinetics toward salicylic acid from aqueous solution. Chem. Eng. J. 2012, 191, 195−201. (34) Yousef, R. I.; El-Eswed, B.; Al-Muhtaseb, A. H. Adsorption characteristics of natural zeolites as solid adsorbents for phenol removal from aqueous solutions: kinetics, mechanism, and thermodynamics studies. Chem. Eng. J. 2011, 171, 1143−1149. (35) Duran, C.; Ozdes, D.; Gundogdu, A.; Senturk, H. B. Kinetics and isotherm analysis of basic dyes adsorption onto almond shell (Prunus dulcis) as a low cost adsorbent. J. Chem. Eng. Data 2001, 5, 2136−2147. (36) Wu, C. H. Adsorption of reactive dye onto carbon nanotubes: equilibrium, kinetics and thermodynamics. J. Hazard. Mater. 2007, 144, 93−100.

G

DOI: 10.1021/acs.jced.6b00174 J. Chem. Eng. Data XXXX, XXX, XXX−XXX