You are on page 1of 40

Accepted Manuscript

Title: Formosa papaya seed powder (FPSP): Preparation,


characterization and application as an alternative adsorbent for
the removal of crystal violet from aqueous phase

Author: Flávio A. Pavan Eveline S. Camacho Eder C. Lima


Guilherme L. Dotto Vivian T.A. Branco Silvio L.P. Dias

PII: S2213-3437(13)00264-9
DOI: http://dx.doi.org/doi:10.1016/j.jece.2013.12.017
Reference: JECE 242

To appear in:

Received date: 7-10-2013


Revised date: 30-11-2013
Accepted date: 21-12-2013

Please cite this article as: F.A. Pavan, E.S. Camacho, E.C. Lima, G.L. Dotto,
V.T.A. Branco, S.L.P. Dias, Formosa papaya seed powder (FPSP): Preparation,
characterization and application as an alternative adsorbent for the removal of crystal
violet from aqueous phase, Journal of Environmental Chemical Engineering (2013),
http://dx.doi.org/10.1016/j.jece.2013.12.017

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
1 Formosa papaya seed powder (FPSP): Preparation, characterization and application as

2 an alternative adsorbent for the removal of crystal violet from aqueous phase

4 Flávio A. Pavana*, Eveline S. Camachoc , Eder C. Limab, Guilherme L. Dottod, Vivian T. A.

t
5 Brancoa, Silvio L. P. Diasb

ip
6

cr
a
7 Institute of Chemistry, Federal University of Pampa, UNIPAMPA, Bagé, RS, Brazil.
b
8 Institute of Chemistry, Federal University of Rio Grande do Sul, UFRGS, Porto Alegre, RS,

us
9 Brazil.
c

an
10 Institute of Chemistry, Federal University of Pelotas, UFPel, Pelotas, RS, Brazil.
d
11 Chemical Engineering Department, Federal University of Santa Maria-UFSM, Santa Maria,
12
M
13 RS, Brazil.

14
d
15
te

16

17
p

18
ce

19

20
Ac

21

22

23

24

25
26 *Corresponding author: Flávio A. Pavan Institute of Chemistry, UNIPAMPA 96412-420 Bagé, RS, Brazil

27 Phone: +55 53 32472367 E-mail address: flavio.pavan@unipampa.edu.br or flavioapavan@yahoo.com.br

1
Page 1 of 39
28 Abstract

29

30 Formosa papaya (Carica papaya L.) seed powder (FPSP), a solid byproduct from

31 industrial and agricultural activities, was used as an alternative adsorbent for the removal of

t
32 crystal violet (CV) from aqueous solutions. The FPSP was characterized by specific surface

ip
33 area (BET), scanning electron microscopy (SEM), infrared spectroscopy (FT–IR), thermal

cr
34 analysis (TGA) and Boehm titration techniques. The effects of initial pH of solution,

35 adsorbent dosage, contact time and initial dye concentration on CV adsorption were studied

us
36 using batch contact mode at 25 °C. Kinetic data were evaluated by pseudo–first order,

an
37 pseudo–second order and Elovich models. The equilibrium adsorption was analyzed by

38 Langmuir, Freundlich and Redlich-Peterson isotherms. Desorption studies were also


M
39 performed. The results indicated that the pseudo–second order model agreed very well with

40 the kinetic data. The adsorption of CV onto FPSP was well fitted using Langmuir isotherm.
d

41 The maximum adsorption capacity obtained by the Langmuir model was 85.99 mg g–1.
te

42 Regeneration of FPSP adsorbent was obtained satisfactory using 1.00 mol L–1 CH3COOH as

43 eluent. These results demonstrated that FPSP is a promising adsorbent to remove CV from
p

44 aqueous solutions.
ce

45

46 Keywords: Adsorption; Aqueous phase; Characterization techniques; Crystal violet; Formosa


Ac

47 papaya seed.

48

49 1. Introduction

50

51 The contamination of water caused by dye molecules is one of the greatest problems of

52 the contemporaneous society. Due to the rapid industrial growth of textile industries in the

2
Page 2 of 39
53 world, large volumes of dye and colored compounds are generated annually in the residual

54 effluents. It is known that dye molecules are undesirable to the aquatic environment due to

55 their high toxicity and poor degradability [1,2]. Another important question is the aesthetic

56 pollution, which is caused by the dye molecules even at low concentrations [3]. Thus, it is

t
57 necessary to remove such contaminants from the industrial effluents, in order to protect the

ip
58 environment and offer people a better quality of live. In this way, several scientific studies

cr
59 have shown that popular processes such as electrochemical [4], biological [5], liquid–liquid

60 extraction [6], oxidation [7] and adsorption [8] can be used with success for the removal of

us
61 different classes of dyes from aqueous phase. Among these, adsorption has been extensively

an
62 employed for the removal of dye molecules from different types of water due to its easy

63 operational conditions [9]. Activated carbon is one of the most efficient adsorbents used in
M
64 adsorption processes; however, its application in industrial scale, mainly in developing

65 countries, is limited due to its high cost of production [10].


d

66 The scientific community has focused on searching for cheaper and more efficient
te

67 materials in order to substitute the activated carbon. Among the promising materials are the

68 materials originated from vegetal biomass (leaves, root, wood, seeds, etc) [11]. These vegetal
p

69 adsorbents have important characteristics that are desirable for the adsorption process such as
ce

70 good adsorption capacity, possibility of regeneration and reuse, and no pollution of the

71 environment. Furthermore, it can be used in large industrial scale being very important from
Ac

72 economic viewpoint. Several papers have been published in the current literature showing the

73 applicability of various different vegetal adsorbents for the removal of dye from water,

74 including, yellow passion fruit peel powder [12], lotus leaf [13], pine leaves [14], olive

75 pomace [15], Parthenium hysterophorus [16], tomato plant root [17] and others. Nevertheless,

76 studies involving the use of papaya seed for the removal of dye from aqueous phase still are

77 scarce. Indeed, only two works have been published using papaya seed as adsorbent for

3
Page 3 of 39
78 removal of dye from aqueous solutions until now: one reported by Hameed [18] and other by

79 Unuabonah [19].

80 Papaya (Carica Papaya L.) is a native fruit of tropical America and widely distributed

81 in all areas of the tropical world, where it is produced mainly for the consumption of fresh

t
82 fruit, juice, or jams. Papaya fruit (Carica papaya, L) is the most popular fruit in Brazil, which

ip
83 is currently the world’s largest producer, responding to a quarter of the global output and the

cr
84 third largest exporter [20]. Two cultivars of Carica Papaya L. are produced in Brazil, namely

85 the Papaya Papaya and the Formosa Papaya. Due to the high annual consumption, great

us
86 amounts of papaya seed are generated in the industrialization process and are considered

an
87 wastes. Papaya seed is composed basically of proteins, crude fibers, ashes, lipids,

88 carbohydrates and organic acids [21]. This variety of biomolecules indicates that some
M
89 functional groups exist on its structure. These groups can be capable to bind with the dyes.

90 The aim of the present study is to report the use of Formosa Papaya Seed Powder
d

91 (FPSP), a solid agricultural waste, as an alternative adsorbent for the removal of hazardous
te

92 crystal violet (CV), a cationic dye considered a mutagen and mitotic poison. Firstly, Formosa

93 papaya seed powder was characterized by Fourier transform infrared (FT–IR) spectroscopy,
p

94 Boehm titration, scanning electron microscopy (SEM), nitrogen adsorption–desorption (BET)


ce

95 isotherm, and thermogravimetric (TGA) techniques. After, for adsorption studies, the effects

96 of initial pH of solution (2.0–10.0), adsorbent dosage (2.0–20.0 g L–1), contact time (5.0–
Ac

97 300.0 min) and initial dye concentration (10.0–40.0 g L–1) were investigated using batch mode

98 at temperature 25°C. Then, kinetic and equilibrium studies were performed through the fit of

99 usual models. Finally, desorption was evaluated using different concentrations of CH3COOH

100 (0.05–2.00 mol L–1) and the possible interactions FPSP–CV were identified.

101

4
Page 4 of 39
102 2. Materials and Methods

103

104 2.1. Solution and reagents

105

t
106 The CV cationic dye (C.I. number 42555; Basic Violet 3 BV 10; molecular formula

ip
107 C25H30N3Cl, molecular weight 407.99 g mol–1) was obtained from Sigma Chemical Co., USA,

cr
108 with analytical grade. The CV chemical structure is presented in Fig. 1.

109 The stock solution was prepared by dissolving accurately the weighed dye in de–

us
110 ionized water to obtain the concentration of 2000.0 mg L-1. Working solutions were obtained

an
111 by diluting the dye stock solution to the required concentrations. De–ionized water was used

112 throughout for solution preparations. The initial pH of each solution was adjusted with 0.10

mol L–1 NaOH and HCl solutions using a pH–meter (Digimed, DM 20, Brazil) for the
M
113

114 measurements.
d

115
te

116 2.2 Adsorbent preparation

117
p

118 Formosa Papaya fruits (Carica papaya L) were purchased at local market. The seeds
ce

119 were removed manually from fresh papaya fruit and washed (more than 5 times) with distilled

120 water to remove dust and other impurities. Afterwards, the seeds were dried at 60 °C for 48 h
Ac

121 (Sterilifer, SX450A, Brazil), crushed (Quimis, Q298A21, Brazil) and sieved through a 250

122 μm sieve. Subsequently, the seeds were washed with doubly distilled water and then dried in

123 an oven at 60 °C for 12 h (Sterilifer, SX450A, Brazil). The powdered seeds obtained were

124 then denoted Formosa Papaya Seed Powder (FPSP).

125

126

5
Page 5 of 39
127 2.3 Adsorbent characterization

128

129 The specific surface area and porosity of FPSP were determined by the Brunauer,

130 Emmett and Teller (BET) and Barret Joyner and Halenda (BJH) methods by the N2 adsorption

t
131 technique using a surface area analyzer (Quantasorb, QS–7, USA). The FPSP was previously

ip
132 degassed at 120 °C in vacuum for 3 h. N2 analysis was carried out using a bath at temperature

cr
133 of –196.0 °C.

us
134 The point zero charge (pHpzc) of the FPSP was determined firstly by pH measurement

135 technique using the following procedure: 50.0 mL of 0.10 mol L–1 NaCl solution was placed

an
136 in a closed 100.0 mL capped Erlenmeyer flask. The pH of each solution was adjusted to

137 values of 2.0 to 10.0. An amount of 0.200 g of FPSP sample was added to the flask and the
M
138 final pH (pHf) measured after 48 h under of stirring at 25 °C. The pHpzc is the point where the

139 curve pH final (pHf) versus pH initial (pHi) crosses the line and equals to pH final [22]. In
d

140 addition, the pHpzc of FPSP also was determined by using a Zetasizer (Malvern, Nano –
te

141 ZS,USA).

142 The concentrations of acid groups (carboxylic, phenol and lactonic) present onto FPSP
p

143 were determined by Boehm titration method [23].


ce

144 Analyses of moisture, ashes, fiber, pH, and specific density were carried out according

to the methodology of Association Official of Analytical Chemistry [24].


Ac

145

146 The information of the chemical structure of FPSP was obtained with the help of

147 Fourier transform infrared (FT–IR) spectroscopy (Shimadzu, 8300, Japan). Amounts of 50.0

148 mg of the sample were pressed with dried potassium bromide. The spectra were obtained with

149 a resolution of 4 cm−1 with 100 cumulative scans over the range of 4000–1000 cm–1. Analyses

150 were carried–out at 25 °C in a dehumidified room.

151 The investigation of surface morphology of non–loaded (native) and dye–loaded FPSP

6
Page 6 of 39
152 was made by scanning electron microscopy (SEM) (Jeol, JSM 5800, Japan) with working

153 voltage of 10 kV at magnification of 1500 times.

154 Thermogravimetric (TGA) and derivative thermogravimetric (DTG) curves of FPSP

155 were carried out by using a thermogravimetric analyzer TGA (Q5000IR, TA instrument,

t
156 USA) under experimental conditions: initial temperature 50 °C; final temperature: 700 °C;

ip
157 heating rate: 10 °C min–1 and nitrogen flow rate 25 mL min–1.

cr
158

us
159 2.4 Batch contact adsorption experiments

160

an
161 The effects of various parameters on CV adsorption by FPSP were investigated using

162 batch contact adsorption at 25 °C (Quimis, G225M, Brazil). These experiments were
M
163 performed with 50.0 mL of CV solution (initial concentration 50.0 mg L–1), at different

164 contact times (5.0–300.0 min), adsorbent dosage (2.0–20.0 g L–1) and pH values (2.0–10.0).
d

165 For adsorption equilibrium studies 0.6 g adsorbent were placed in a series of conical flasks
te

166 (125.0 mL) each one containing 50.0 mL of different initial dye concentrations (5.0–1000.0

167 mg L–1) at pH 8.0. The conical flasks were shaken in a rotary orbital shaker at 150 rpm for
p

168 60.0 min (Quimis, G225M, Brazil). The adsorbent was separated from the liquid phase by
ce

169 centrifugation (Quimis, Q222T216, Brazil) and the remaining dye in the solution was then

determined by using a UV–Vis spectrophotometer (Varian, Cary 50 Bio, USA) with 1.0 cm
Ac

170

171 path length cell. Absorbance measurements were made at 584.0 nm. Dye solutions with

172 absorbance measurements higher than 1.0 were diluted. The experiments were carried out in

173 replicate (n=3) and blanks were performed. The amount of dye adsorbed per gram of

174 adsorbate at equilibrium (qe), at any time (qt) and the dye removal percentage (R) were

175 obtained by the Eqs. (1–3), [12, 14], respectively:

176

7
Page 7 of 39
C0 − Ce
177 qe = V (1)
m

178

C0 − C t
179 qt = V (2)
m

t
ip
180

C0 − Ce
R=

cr
181 100 (3)
C0

us
182

183 where, Co is the initial CV concentration (mg L–1), Ce is the equilibrium CV concentration

an
184 (mg L–1), Ct is the CV concentration at any time (mg L–1), m is the adsorbent amount (g) and

185 V is the volume of dye solution (L).


M
186

187 2.5 Kinetic and equilibrium models


d

188
te

189 The successful adsorption process depends of the kinetic parameters. Knowing the

190 adsorption kinetics it is possible to devise and conduct the process more efficiently. To
p

191 establish the necessary equilibrium time for maximum CV adsorption on the FPSP, three
ce

192 different kinetic models, namely pseudo–first order [25], pseudo–second order [26] and

193 Elovich [27] were fitted with the experimental data. Furthermore, it is known that equilibrium
Ac

194 isotherms are essential to the understanding of the interaction mechanism between dye and

195 adsorbent. To study the relationship between adsorbed CV per unit of adsorbent (mg g–1) and

196 non–adsorbed CV concentration in the aqueous phase (mg L–1) at equilibrium, three

197 equilibrium isotherms nominated Langmuir [28], Freundlich [29] and Redlich–Peterson [30]

198 were used. The kinetic and equilibrium models are presented in Table 1.

8
Page 8 of 39
199 The kinetic and equilibrium parameters were determined by the fits on the models

200 (Table 1) with the experimental data through nonlinear regression using the Quasi–Newton

201 estimation method. The calculations were carried out by the Statistic 7.0 software (Statsoft,

202 USA). The fit quality and the accuracy of the parameters were measured through

t
203 determination coefficient (R2), adjusted determination coefficient (R2adj) and Chi–square test

ip
204 (χ2) [31].

cr
205

us
206 2.6 Batch desorption studies

207

an
208 Batch desorption studies were carried with goal to restore the adsorbent as a function

209 of concentration of CH3COOH, as the eluent. For this propose, 12.0 g L-1 of dye-loaded
M
210 adsorbent obtained from adsorption experiments, was mixed with 50.0 mL of (0.05, 0.10 0.20,

211 0.50 1.00, 1.50 and 2.00) mol L–l CH3COOH solution in the Erlenmeyer flask, the mixture
d

212 was agitated with help of rotator shaker at 150 rpm for 60.0 min to reach equilibrium.
te

213 Afterwards, the samples were centrifuged at 5000.0 rpm for 10.0 min, the dye concentration

214 released on the desorbing solution was determined at the maximum wavelength of CV dye
p

215 (584.0 nm).


ce

216

217 3. Results and discussion


Ac

218

219 3.1 Characterization of the adsorbent

220

221 The BET surface area, total pore volume and average pore diameter of the FPSP were

222 found be 1.38 m2 g-1, 0.0024 cm3 g-1 and 15 Å, respectively. This low value of surface area is

223 characteristic of biomass materials [32,33]. The point zero charge (pHpzc) of the FPSP

9
Page 9 of 39
224 determined by pH measurement technique was 6.851 (Fig. 2(a)). This result was confirmed by

225 Zeta potential experiment (Fig. 2(b)).

226 This result is in agreement with the data reported by Unuabonah et al. [21]. Other

227 characteristics of FPSP adsorbent are summarized in Table 2.

t
228 The infrared spectroscopy technique was used to identify the functional groups on the

ip
229 FPSP which can be responsible for binding dye molecules, and also to elucidate the dye–

cr
230 adsorbent interactions. The FT–IR spectrum of FPSP before adsorption (Fig. 3 (a)) showed an

231 intense absorption band at 3280 cm–1, which is relative to the −OH groups on the FPSP

us
232 surface [12]. The bands at 2920 cm–1 and 2857 cm–1 can be assigned to the stretching of OH

an
233 groups bound to the methyl group (C−OH) [8,34,35]. The band observed at 1740 cm–1 is

234 relative to the carbonyl group (C=O) of carboxylic acid or ester [12]. The band at 1638 cm–1
M
235 can be attributed to the −C=O stretching of carboxylic acid with intermolecular hydrogen

236 bond [12,36]. The adsorption bands at 1460 cm–1 and 1315 cm–1could be attributed to the
d

237 −C−O group [8]. The bands around 1235 cm–1 and 1160 cm–1 can be assigned to ether, ester or
te

238 phenol groups [36]. In brief, the FT–IR analysis indicated the presence of functional groups

239 such as OH, C=O and C−O on the surface of FPSP. These data were also confirmed by
p

conventional Boehm titration, where quantitative values of phenolic, carboxylic and carbonyl
ce

240

241 groups were measured (see Table 2). From the FT–IR spectrum of FPSP adsorbed with CV
Ac

242 (Fig. 3 (b)), it was possible to observe some changes in relation to the native FPSP spectrum.

243 After the adsorption process a reduction of sharpness of the bands at 2920 cm–1 and 2853 cm–1

244 was observed, indicating that the acid hydroxyl groups (C−OH) on the FPSP surface

245 participated of the interaction with CV molecules. The bands at 1740 cm–1 and 1638 cm–1

246 shifted to 1732 cm–1 and 1653 cm–1, respectively. Also the band at 1235 cm–1 shifted to 1227

247 cm–1. In addition, a decrease in the intensities of these bands was observed after adsorption.

248 Another change was the appearance of new strong bands at 1582 cm–1, which can be

10
Page 10 of 39
249 attributed to the C–N group of dye molecules (See Fig. 1) [35,36]. These results show that the

250 CV adsorption onto FPSP occurred satisfactorily, and that, the functional groups such as OH,

251 C=O, and C−O participate of the interaction between CV and adsorbent. These results are

252 concordance with the studies of Akar et al. [34] for the decolorization of Reactive blue 49 by

t
253 capsicum annuum seed, and Lin et al. [35] for the removal of CV using powdered mycelial

ip
254 biomass of Ceriporia lacerata.

cr
255 Surface morphologies of native and dye–loaded FPSP were investigated by SEM. The

SEM images of FPSP and dye–loaded FPSP are shown in Fig. 4 (a) and 4 (b), respectively.

us
256

257 Results show that the change in textural properties of FPSP after CV adsorption (Fig 4(b)) is

an
258 due to the accumulation of dye onto the adsorbent surface indicating that the adsorption

259 phenomena occurred. Similar results were obtained by Kumar and Ahmad [36] for the
M
260 removal of CV using treated ginger waste as biosorbent.

261 Thermogravimetric analysis (TGA) was used to investigate the thermal profile of the
d

262 FPSP adsorbent. The TGA curve of FPSP adsorbent was made in N2 atmosphere and is
te

263 presented in Fig. 5. The modifications on the FPSP structure appeared between 60 °C and

264 150°C and were intensified between 250 °C and 450 °C. The initial weight loss (60–150 °C)
p

265 was attributed to the desorption of water molecules from the surface and pores of FPSP. The
ce

266 weight loss in the initial steps was about 4%. The second weight loss (61%) between 250 °C
Ac

267 and 450 °C could be ascribed to the degradation of hemicelluloses and cellulose [37]. The

268 weight loss (29%) between 450 °C and 700 °C was attributed to the degradation of lignin

269 [37].

270

271 3.2 Effects of initial pH of solution and adsorbent dosage

272

11
Page 11 of 39
273 Fig. 6 (a) shows the effect of the initial pH of solution on the adsorption capacity of

274 CV on FPSP. In Fig. 6 (a), it is possible to note that when the pH values increased from 2.0 to

275 8.0, an increase on the CV removal percentage occurred (41.3 to 97.3%). In the pH range 8.0–

276 10.0 the CV removal percentage remained nearly constant (variation observed lower that

t
277 3.0%) (Fig. 6 (a)). This shows that the adsorption of the CV cationic dye onto FPSP was

ip
278 favored at alkaline pH region (pH 8.0–10.0). The point zero charge (pHpzc) of FPSP could be

cr
279 used to explain this influence. In this study the pHpzc of FPSP was 6.85. So, at pH values

280 lower than 6.85 (pH<pHpzc), the surface of FPSP was positively charged, inhibiting the

us
281 adsorption of cationic dye due to the electrostatic repulsion between cationic structure of CV

an
282 and FPSP. On the other hand, at pH values above 6.85 (pH>pHpzc) the surface of FPSP was

283 negatively charged, facilitating the electrostatic attraction between the cationic dye, and
M
284 consequently, increasing the CV removal percentage. Similar results were obtained by Jain

285 and Jayaram [38] using wood apple shell as biosorbent for the removal of CV and also by
d

286 Saeed et al. [39] using grapefruit peel as adsorbent for the removal of CV. Based on the above
te

287 mentioned reasons, the initial pH of 8.0 was selected as the more adequate value for the other

288 adsorption experiments.


p

289 The adsorbent dosage effect on the CV removal percentage is shown in Fig. 6 (b). It
ce

290 was found that the CV removal percentage increased from 41.0 to 98.4% when the FPSP

291 dosage was increased from 1.0 to 12.0 g L–1. This fact is due to the increase in surface area
Ac

292 and number of active sites to bind the dye. Above 12.0 g L–1, the CV removal percentage was

293 constant. This behavior may be due the agglomeration of adsorbent particles. These results are

294 in agreement with the literature about adsorption of dyes using low cost adsorbents [40–42].

295 Based on these results, it was found that 12.0 g L–1 is an adequate adsorbent dosage for the

296 removal CV from aqueous solutions, and so, was used for subsequent experiments.

297

12
Page 12 of 39
298 3.3 Effects of contact time and initial dye concentration

299

300 The contact time effect was studied at pH 8.0, temperature of 25 °C, adsorbent dosage

301 of 12.0 g L–1, under different initial dye concentrations (from 10.0 to 40.0 mg L–1). The

t
302 contact time effect is shown in Fig. 7 (a). It is possible to observe in Fig. 7 (a), that more than

ip
303 96.0% of saturation was achieved within 60.0 min, while only 3.0% was adsorbed until 300.0

cr
304 min. This occurred because at the initial adsorption stages, FPSP has a large available surface

305 area, compared with the concentration of dye molecules, favoring the high adsorption rate.

us
306 However, after the initial adsorption stages (above 60.0 min) a competition occurred between

an
307 dye molecules for the adsorption sites on the surface of FPSP, resulting in a decrease of

308 adsorption rate. With respect to the initial dye concentration effect, it was found in Fig. 7 (a),

that the amount of CV adsorbed increased (from 1.81 to 9.36 mg g–1) with the increase of the
M
309

310 initial dye concentration (10.0 to 40.0 mg L–1). This occurred because at higher initial
d

311 concentrations, the concentration gradient between the bulk solution and the adsorbent surface
te

312 is higher, facilitating the external mass transfer [43]. In parallel, the internal mass transfer is

313 facilitated at high values of initial dye concentration [44]. Similar trend was reported by
p

314 Naveen et al. [45].


ce

315

316 3.4 Kinetic studies


Ac

317

318 Kinetic studies were carried out at pH 8.0, temperature of 25 °C, adsorbent dosage of

319 12.0 g L–1, under different initial dye concentrations (from 10.0 to 40.0 mg L–1). The models,

320 namely pseudo–first order, pseudo–second order and Elovich (Table 1) were fitted with the

321 experimental data. In the case of pseudo–first order and pseudo–second order equations, kf

322 (min–1) and ks (g mg–1 min–1) are the pseudo–first order and pseudo–second order rate

13
Page 13 of 39
323 constants, respectively, and qe is theoretical value of adsorption capacity for each model. In

324 the Elovich equation, α (mg g–1 min–1) is the initial adsorption rate and β (g mg–1) is

325 desorption constant. The experimental kinetic curves with the respective fits of the models are

326 shown in Figs. 7 (b), 7 (c) and 7 (d). The kinetic parameters and the values of R2, adjusted,

t
327 R2adj and χ2 are shown in Table 3.

ip
328 Based on the high values of R2, R2adj and low values of χ2 presented in Table 3, it can

cr
329 be affirmed that the pseudo–second order model was the more adequate to represent the

us
330 adsorption of CV on FPSP. The pseudo–second order model was also adequate for the

331 adsorption of CV onto Ceriporia lacerate [35] and grapefruit peel [39]. As shown in Table 3,

an
332 the qe values (pseudo–second order model) increased with the increase of the initial dye

333 concentration, while ks values were decreased. These results indicated that when the amount
M
334 of dye–adsorbed on the FPSP surface was increased, the diffusion process was hindered and a

335 high competition of dye molecules for fixed functional groups occurred. In addition to the
d

336 above discussed parameters, the initial sorption rate (h0) was calculated by the Eq. (4) [26]:
te

337

h 0 = k s (q e )
2
338 (4)
p
ce

339

340 Table 3 shows that the increase in initial dye concentration caused an increase in the
Ac

341 initial sorption rate (h0), indicating that, in the initial stages, the adsorption capacity and the

342 adsorption rate were higher at 40 mg L–1.

343

344 3.5 Equilibrium studies

345

346 The experimental equilibrium curve (Fig. 8) was obtained at pH 8.0, temperature of 25

347 °C and adsorbent dosage of 12.0 g L–1. It was found in Fig. 8 that the adsorption isotherm

14
Page 14 of 39
348 curve was characterized by an initial step with increase in adsorption capacity followed by a

349 convex shape. The initial step indicates a great FPSP–CV affinity and numerous readily

350 accessible sites. The convex shape suggests the formation of a monomolecular layer of the CV

351 dye on the FPSP surface [46].

t
352 Langmuir, Freundlich and Redlich–Peterson models (Table 1) were used to fit the

ip
353 experimental data. The Langmuir isotherm implies a monolayer adsorption on the

cr
354 homogeneous surface. The Langmuir isotherm contains two parameters: Qmax (mg g–1),

355 representing the maximum amount of the dye adsorbed at equilibrium and KL (L mg–1), the

us
356 Langmuir constant [28]. The Freundlich equation is an empirical model, which considers that

an
357 the adsorption occurs on a heterogeneous surface and the active sites have different energies.

358 The Freundlich equation is represented by two parameters: KF (mg g–1 (mg L–1)–1/nF) is the
M
359 Freundlich constant and nF is relative to the adsorption intensity [29]. Redlich–Peterson is

360 another isotherm model widely used to characterize adsorption processes over a wide range of
d

361 adsorbate concentration. The Redlich–Peterson equation contains three parameters: KRP (L g–
te

1
362 ) is the Redlich–Peterson constant, aRP (mg L–1) affinity coefficient and g is the heterogeneity

363 parameter. The g parameter assumes values between 0 and 1 [30]. The isotherm parameters,
p

364 R2, R2adj and χ2 are presented in Table 4.


ce

365 The high values of R2, R2adj and low values of χ2 (Table 4) demonstrated that the

366 Langmuir model was the more adequate to represent the adsorption of CV on FPSP. Based on
Ac

367 the Langmuir isotherm, the maximum adsorption capacity obtained was 85.99 mg of CV per

368 gram of FPSP. To show the potentially of FPSP adsorbent prepared in the present study, a

369 comparison with other adsorbents employed for the removal of CV was realized (Table 5).

370 Comparing the maximum adsorption capacities of the different adsorbents with the FPSP

371 adsorption capacity we can conclude that it is an effective adsorbent.

372

15
Page 15 of 39
373 3.6 Desorption

374

375 Desorption studies were performed to evaluate the possibility of recovery CV, as well

376 as, regeneration of the adsorbent. Fig. 9 shows the CV desorption percentage as a function of

t
377 the eluent (CH3COOH) concentration (0.05–2.00 mol L–1). The best desorption was obtained

ip
378 with 1.00 mol L–1 of CH3COOH, with CV recovery of 96.5%. Similar results were obtained in

cr
379 the CV adsorption on Coniferous pinus bark powder [47] and Ricinus cummunis pericarp

380 carbon as adsorbents [53]. It was also verified that five adsorption–desorption cycles were

us
381 possible, with losses of 6.0% in the adsorption capacity. Furthermore, if the CV adsorption

an
382 onto FPSP occurred in basic solution (as above demonstrated in section 3.2), and the

383 desorption occurred effectively in acid media, it can be inferred that electrostatic interactions
M
384 occurred between CV and FPSP [54].

385
d

386 3.7 Interactions FPSP–CV


te

387

388 Based on the FT–IR spectral analysis and others results previously discussed, the
p

389 probable mechanisms for CV adsorption from aqueous solution onto FPSP are discussed.
ce

390 The FT-IR analysis indicate that OH and COOH groups are effective in the adsorption of CV

391 from aqueous solution. For basic pH (selected pH =8.0) the all hydroxyl and carboxyl groups
Ac

392 are deprotonated so, this means that the FPSP surface is negatively charged and positively

393 charged dyes molecules can interact strongly with O- and COO- sites present onto adsorbent.

394 This result indicate that the cationic dye molecules are adsorbed onto FPSP by electrostatic

395 attraction. The electrostatic attraction between FPSP and CV molecules is presented below:

396

397 For hydroxyl it is:

16
Page 16 of 39
398

399 FPSP − OH + CV − Cl → FPSP − O − CV + + H + Cl −

400

401 For carboxylic acid

t
402

ip
403 FPSP − COOH + CV − Cl → FPSP − COO − CV + + H + Cl −

cr
404

405 Others possible adsorption mechanisms between FPSP and CV molecules also may

us
406 be attributed to (1) van der Waals interaction between hydrophobic parts of CV and FPSP

407 and (2) hydrogen bonding interaction between nitrogen of CV molecules and FPSP surface.

408
an
M
409 4. Conclusions

410

In this research, Formosa papaya (Carica papaya L.) seed powder (FPSP) was
d
411

prepared, characterized and used as an alternative adsorbent for the removal of crystal violet
te

412

413 (CV) from aqueous solutions. The FPSP presented specific surface area of 1.38 m2 g–1, pHpzc
p

414 of 6.85, a rough, porous and heterogeneous surface with O−H, C=O and C−O functional
ce

415 groups and its thermal degradation occurred close to 400 °C. Regarding the adsorption

416 process, the more adequate conditions for CV adsorption were: pH from 8.0 to 10.0 and
Ac

417 adsorbent dosage of 12.0 g L–1. The minimum contact time required to reach the equilibrium

418 was 60.0 min. The adsorption of CV on FPSP followed a pseudo–second order kinetic model.

419 The Langmuir model presented best fit with the equilibrium experimental data, being the

420 maximum adsorption capacity of 85.99 mg g–1, obtained at 25 °C. Five adsorption–desorption

421 cycles were possible using 1.00 mol L–1 of CH3COOH. Electrostatic interactions occurred

422 between FPSP and CV under basic conditions. Based on our results it is possible to confirm

17
Page 17 of 39
423 that FPSP, a bio–solid waste, is a good option for the removal of CV from aqueous solutions,

424 since it is considered cheap, available in large amounts and effective.

425

426 Acknowledgements

t
427

ip
428 The authors are grateful to the CNPq, FAPERGS and CAPES for financial support.

cr
429 We are also grateful to Center of Electron Microscopy (CME–UFRGS) for the use of the

430 SEM microscope.

us
431

an
432

433
434 References
M
435

436 [1] Y.M. Kolekar, S.P. Pawar, K.R. Gawai, P.D. Lokhande, , Y.S. Schouche, K.M. Kodam,
d

437 Decolorization and Degradation of Disperse Blue 79 and Acid Orange 10 by Bacillus
te

438 fusiformis KMK5 Isolated from the Textile Dye Contaminated Soil, Bioresour. Technol. 99
p

439 (2008) 8999.


ce

440 [2] M.K. Purkait, A. Maiti, S. das Gupta, S. De, Removal of congo red using activated carbon

441 and its regeneration, J. Hazard. Mater. 145 (2007) 287.


Ac

442 [3] T. Robinson, G. McMullan, R. Marchant, P. Nigam, Remediation of dyes in textile

443 effluents: a critical review on current treatment technology with a proposed alternative,

444 Bioresour. Technol. 77 (2001) 247.

445 [4] C.L. Yang, J. McGarrahan, Electrochemical coagulation for textile effluent decolorization,

446 J. Hazard. Mater. 127 (2005) 40.

447 [5] M. Isik, D.T. Sponza, Anaerobic/Aerobic Treatment of a SimulatedTextile Wastewater,

448 Sep. Purif. Technol. 60 (2008) 64.

18
Page 18 of 39
449 [6] A.S. Mahmoud, A.E. Ghaly, M. S. Brooks, Removal of Dye From Textile Wastewater

450 Using Plant Oils under Different pH and Temperature Conditions, Am. J. Environ. Sci. 3

451 (2007) 205.

452 [7] D.S. Kim, Y.S. Park, Comparison Study of Dyestuff Wastewater Treatment by the

t
453 Coupled Photocatalytic Oxidation and Biofilm, Process, Chem. Eng. J. 139 (2008) 256.

ip
454 [8] R. Gottipati, S. Mishra, Application of Biowaste (waste generated in biodiesel plant) as an

cr
455 adsorbent for the removal of hazardous dye- methylene blue- from aqueous phase, Brazilian J.

456 Chem. Eng. 27 (2010) 357.

us
457 [9] X.S. Wang, J.P. Chen, Removal of the Azo Dye Congo Red from Aqueous Solutions by

an
458 the Marine Alga Porphyra yezoensis Ueda, Clean - Soil, Air, Water. 37 (2009) 793.

459 [10] D.C.W. Tsang, J. Hu, M.Y. Lui, W. Zang, K.C.K. Lai, I.M.C. Lo, Activated carbon
M
460 produced from waste wood pallets: adsorption of three classes of dyes, Water Air Soil Pollut.

461 184 (2007) 141.


d

462 [11] P. Sharma, H. Kaur, M. Sharma, V. Sahore, A review on applicability of naturally


te

463 available adsorbents for the removal of hazardous dyes from aqueous waste, Environ Monit

464 Assess. 183 (2011) 151.


p

465 [12] F.A. Pavan, E.C. Lima, S.L.P. Dias, A.C. Mazzocato, Methylene blue biosorption from
ce

466 aqueous solutios by yellow passion fruit waste, J. Hazard. Mater, 150 (2008) 703.

467 [13] X. Han, W. Wang, X. Ma, Adsorption characteristics of methylene blue onto low cost
Ac

468 biomass materiallotus leaf, Chem. Eng. J. 171 (2011) 1.

469 [14] F. Deniz, S. Karaman, Removal of an azo-metal complex textile dye from colored

470 aqueous solutions using an agro-residue, Microchem. J. 99 (2011) 296.

471 [15] T. Akar, I. Tosun, Z. Kaynak, E. Ozkara, O. Yeni, E. N. Sahin, Akar, S.T. An attractive

472 agro-industrial by-product in environmental cleanup: Dye biosorption potential of untreated

473 olive pomace, J. Hazard. Mater. 166 (2009) 1217.

19
Page 19 of 39
474 [16] H. Lata, V.K. Garg, R.K. Gupta, Removal of a basic dye from aqueous solution by

475 adsorption using Parthenium hysterophorus: An agricultural waste, Dyes Pigm. 74 (2007)

476 653.

477 [17] C. Kannan, N. Buvaneswari, T. Palvannan, Removal of plant poisoning dyes by

t
478 adsorption on tomato plant root and green carbon from aqueous solution and its recovery,

ip
479 Desalination 249 (2009) 1132.

cr
480 [18] B.H. Hameed, Evaluation of papaya seeds as a novel non-conventional low cost

481 adsorbent for removal of methylene blue, J. Hazard. Mater. 162 (2009) 939.

us
482 [19] E.I. Unuabonah, G.U. Adie, L.O. Onah, O.G. Adeyemi, Multistage optimization of the

an
483 adsorption of methylene blue dye onto defatted Carica papaya seeds, Chem. Eng. J. 155

484 (2009) 567.


M
485 [20] G.G. Silva, R.G. Diniz, M.E. Silva, Comparing the chemical composition of papaya

486 fruits (Carica papaya L.) in two stages of ripening, Revista Capixaba de Ciência e
d

487 Tecnologia. 3 (2007) 1.


te

488 [21] F.L. Alves, J.M.S Balbino, F.C. Bareto, A cultura do mamoeiro: tecnologia de

489 produção, Vitória, BA, Incaper, (2003) 497.


p

490 [22] M. Khormaei, Nasernejad, B.; Edrisi, M.; Eslamzadeh, T., Copper Biosorption from
ce

491 aqueous solutions by sour orange residue, J. Hazard. Mater. 149 (2007) 269.

492 [23] H.P. Boehm, Some aspects of the surface chemistry of carbon blacks and other carbons,
Ac

493 Carbon, 32 (1994) 759.

494 [24] Association of Official Analytical Chemistry. Official methods of analysis. 10. ed.

495 Washington, 1965.

496 [25] S. Lagergren, About the theory of so-called adsorption of soluble substance, Kungliga

497 Suensk Vetenskapsakademiens Handlingar, 241 (1898) 1.

20
Page 20 of 39
498 [26] G. Mckay, Y.S. Ho, Pseudo-second-order model for sorption processes, Process.

499 Biochem. 34 (1999) 451.

500 [27] H.A. Harouna-Oumarou, H. Fauduet, C. Porte, Y.S. Ho, Comparison of kinetic models

501 for the aqueous solid-liquid extraction of Tilia sapwood in a continuous stirred tank reactor,

t
502 Chem. Eng. Commun. 194 (2007) 537.

ip
503 [28] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum, J.

cr
504 Am. Chem. Soc. 40 (1918) 1361.

505 [29] H.M.F. Freundlich, Über die adsorption on lösungen, Z. Phys.Chem. 57 (1906) 385.

us
506 [30] O. Redlich, D.L. Peterson, A useful adsorption isotherm. J. Phys.Chem. 63 (1959) 1024.

an
507 [31] M.I. El–Khaiary, G.F. Malash, Common data analysis errors in batch adsorption studies,

508 Hydrometallurgy 105 (2011) 314–320.


M
509 [32] M. Antunes, V.I. Esteves, R. Guégan, J.S. Crespo, A.N. Fernandes, M. Giovanela,

510 Removal of diclofenac sodium from aqueous solution by Isabel grape bagasse, Chem. Eng. J.
d

511 192 (2012) 114.


te

512 [33] P. Kumari, P. Sharma, S. Srivastava, M.M. Srivastava, Biosorption studies on shelled

513 Moringa oleifera Lamarck seed powder: Removal and recovery of arsenic from aqueous
p

514 system, Int. J. Miner. Process. 78 (2006) 131.


ce

515 [34] S.T. Akar, A. Gorgulu, T. Akar, S. Celik, Decolorization of Reactive Blue 49 contamined

516 solution by Capsicum annuum seed, Chem. Eng. J. 168 (2011) 125.
Ac

517 [35] Y. Lin, X. He, G. Han, Q. Tian, W. Hu, Removal of Crystal Violet from aqueous solution

518 using powdered mycelial biomass of Ceriporia lacerata P2, J. Environ. Sci- China, 23 (2011)

519 2055.

520 [36] R. Kumar, R. Ahmad, Biosorption of hazardous crystal violet dye from aqueous solution

521 onto treated ginger waste (TGW), Desalination, 265 (2011) 112.

21
Page 21 of 39
522 [37] V. Sricharoenchaikul, D. Atong, Thermal decomposition study on Jatropha curcas L.

523 waste using TGA and fixed bed reactor, J. Anal. Appl. Pyrolysis, 85 (2009) 155.

524 [38] S. Jain, R.V. Jayaram, Removal of basic dyes from aqueous solution by low-cost

525 adsorbent: Wood apple shell (Feronia acidissima), Desalination, 250 (2010) 921.

t
526 [39] A. Saeed, M. Sharif, M. Iqbala, Application potential of grapefruit peel as dye sorbent:

ip
527 Kinetics, equilibrium and mechanism of crystal violet adsorption, J. Hazard. Mater. 179

cr
528 (2010) 564.

529 [40] F. Deniz, S. Karaman, Removal of Basic Red 46 dye from aqueous solution by pine tree

us
530 leaves, Chem. Eng. J. 170 (2011) 67.

an
531 [41] G.O. El-Sayed, Removal of methylene blue and crystal violet from aqueous solutions by

532 palm kernel fiber, Desalination 272 (2011) 225.


M
533 [42] Z. Yao, L. Wang, J. Qi, Biosorption of methylene blue from aqueous solution using a

534 bioenergy forest waste: Xanthoceras sorbifolia seed coat, Clean - Soil, Air, Water 37 (2009)
d

535 642.
te

536 [43] G.L. Dotto, L.A.A. Pinto, Analysis of mass transfer kinetics in the biosorption of

537 synthetic dyes onto Spirulina platensis nanoparticles, Biochem, Eng. J. 68 (2012) 85.
p

538 [44] G.L. Dotto, L.A.A. Pinto, Adsorption of food dyes acid blue 9 and food yellow 3 onto
ce

539 chitosan: Stirring rate effect in kinetics and mechanism, J. Hazard. Mater. 187 (2011) 164.

540 [45] N. Naveen, P. Saravanan, G. Baskar, S. Renganathan, Equilibrium and kinetic modeling
Ac

541 on the removal of Reactive Red 120 using positively charged Hydrilla verticillata, J. Taiwan

542 Inst. Chem. Eng. 42 (2011) 463.

543 [46] G.L. Dotto, E.C. Lima, L.A.A. Pinto, Biosorption of food dyes onto Spirulina platensis

544 nanoparticles: Equilibrium isotherm and thermodynamic analysis, Bioresour. Technol. 103

545 (2012) 123.

22
Page 22 of 39
546 [47] R. Ahmad, Studies on adsorption of crystal violet dye from aqueous solution onto

547 Coniferous pinus bark powder (CPBP), J. Hazard. Mater. 171 (2009) 767.

548 [48] H. Parab, M. Sudersanan, N. Shenoy, T. Pathare, B. Vaze, Use of Agro-Industrial Wastes

549 for Removal of Basic Dyes from Aqueous Solutions, Clean - Soil, Air, Water 37, (2009)

t
550 963.

ip
551 [49] P. Das Saha, S. Chakraborty, S. Chowdhury, Batch and continuous (fixed-bed column)

cr
552 biosorption of crystal violet by Artocarpus heterophyllus (jackfruit) leaf powder, Colloids

553 And Surfaces B-Biointerfaces 92 (2012) 262

us
554 [50] T. Smitha, S. Thirumalisamy, S. Manonmani, Equilibrium and Kinetics Study of

an
555 Adsorption of Crystal Violet onto the Peel of Cucumis sativa Fruit from Aqueous Solution,

556 E-J. Chem. 9 (2012) 1091.


M
557 [51] C. Kannan, N. Buvaneswari, T. Palvannan, Removal of plant poisoning dyes by

558 adsorption on Tomato Plant Root and green carbon from aqueous solution and its recovery.
d

559 Desalination 249 (2009) 1132.


te

560 [52] Z. Muhammad, Removal of Crystal Violet from Water by Adsorbent Prepared from

561 Turkish Coffee Residue, Tenside Surfactants Detergents 49 (2012) 107.


p

562 [53] S. Madhavakrishnan, K. Manickavasagam, R. Vasanthakumar, K. Rasappan, R.


ce

563 Mohanraj, S. Pattabhi, Adsorption of Crystal Violet Dye from Aqueous Solution Using

564 Ricinus Communis Pericarp Carbon as an Adsorbent, E-J. Chem. 6 (2009) 1109.
Ac

565 [54] H. Patel, R.T. Vashi, Adsorption of Crystal Violet Dye onto Tamarind Seed Powder, E-J.

566 Chem. 7 (2010) 975.

567

568

569

570

23
Page 23 of 39
571 Figure captions

572 Figure 1: Chemical structure of CV dye.

573 Figure 2: Determination of (a) pH zero point of charge (pHpzc) and (b) Zeta potential of FPSP.

574 Figure 3: FT–IR analysis of (a) native and (b) dye–loaded FPSP.

t
575 Figure 4: SEM micrographs of (a) native and (b) dye–loaded FPSP.

ip
576 Figure 5: TGA curve of FPSP in N2.

cr
577 Figure 6: Effects of (a) initial pH of solution and (b) adsorbent dosage.

578 Figure 7: Effect of contact time: (a) Experimental curves for CV adsorption on FSPS; (b) Co =

us
579 10.0 mg L–1; (c) Co = 20.0 mg L–1 and (d) Co = 40.0 mg L –1. (Conditions: adsorbent dosage of

12.0 g L–1; pH = 8.0 and temperature of 25°C).

an
580

581 Figure 8: Equilibrium curves for CV adsorption on FSPS (Conditions: adsorbent dosage of
M
582 12.0 g L–1; pH = 8.0 and temperature of 25°C).

583 Figure 9: Desorption results.


d

584
te

585

586
p

587
ce

588

589
Ac

590

591

592

593

594

595

24
Page 24 of 39
596 Table captions

597 Table 1: Kinetic and isotherm models.

598 Table 2: Characteristics of FPSP adsorbent.

599 Table 3: Kinetic parameters for CV adsorption by FPSP at the more adequate conditions.

t
600 Table 4: Isotherm parameters for CV adsorption by FPSP at the more adequate conditions.

ip
601 Table 5: Comparison of adsorption capacities (mg g–1) for the CV onto different adsorbents.

cr
602

603

us
604

an
605

606
M
607

608
d

609
p te
ce
Ac

25
Page 25 of 39
609 Table 1: Kinetic and isotherm models.

Kinetic models Equation

Pseudo–first order
[ (
q t = q e 1− exp − k f t )]
ks qe2 t

t
Pseudo–second order qt =

ip
1 + qe ks t
1
qt = (α β ) + 1 (t )

cr
Elovich
β β
Isotherm models Equation

us
Q max KL Ce
Langmuir qe =
1 + KL Ce

an
1 nF
Freundlich q e = K F Ce
K RP Ce
M
Redlich–Peterson qe = g 0 ≤ g ≤1
1+ a RP Ce
610
d

611
p te
ce
Ac

26
Page 26 of 39
611 Table 2: Characteristics of FPSP adsorbent.

Characteristic Value*
Carboxylic groups (mEq g–1) 3.71 ± 0.05
Phenolic groups (mEq g–1) 1.95 ± 0.03
Lactonic groups (mEq g–1) 3.20 ± 0.08

t
Moisture content (%) 13.51 ± 0.70

ip
Fiber content (%) 13.83 ± 1.02
Ashes content (%) 8.21 ± 0.30

cr
pH of aqueous solution 7.20 ± 0.05
Specific mass (g cm–3, 25 °C) 0.91 ± 0.05

us
*mean±standard error (n=3).
612
613

an
614

615
M
d
p te
ce
Ac

27
Page 27 of 39
615 Table 3: Kinetic parameters for CV adsorption by FPSP at the more adequate conditions.

Co (mg L–1)
Model
10 ± 0.1 20 ± 0.2 40 ± 0.1
Pseudo–first order
kf (min–1) 0.092 ± 0.003 0.064 ± 0.002 0.117 ± 0.001

t
–1
qe (mg g ) 1.778 ± 0.03 5.513 ± 0.02 9.248 ± 0.15

ip
R2 0.982 0.968 0.955

cr
R2adj 0.980 0.964 0.949
–3 –3
χ2 17.6×10 19.0×10 9.43×10–4

us
Pseudo–second order
ks (g mg–1 min–1) 0.078 ± 0.002 0.020 ± 0.001 0.016 ± 0.001
qe (mg g–1) 1.890 ± 0.03 5.921 ± 0.04 9.746 ± 0.11
h0 (mg g–1 min–1)
R2
10
0.993
an 12
0.974
18
0.974
M
R2adj 0.992 0.971 0.971
χ2 17.4×10–6 13.0×10–4 2.15×10–5
Elovich
d

α (mg g–1 min–1) 3.061 ± 0.02 2.694 ± 0.03 57.831 ± 1.1


te

β (g mg–1) 4.231 ± 0.02 1.120 ± 0.05 0.965 ± 0.02


2
R 0.863 0.886 0.806
p

2
R adj 0.846 0.872 0.782
ce

χ2 7.33×10–5 9.13×10–3 3.4×10–5


*mean±standard error (n=3).
616
617
Ac

618

28
Page 28 of 39
618 Table 4: Isotherm parameters for CV adsorption by FPSP at the more adequate conditions.

Langmuir
Qmax (mg g–1) 85.99 ± 2.3
KL (L mg–1) 0.0015 ± 2.10-4
R2 0.986

t
2
R adj 0.985

ip
2
χ 0.0677
Freundlich

cr
–1 –1 –1/nF
KF (mg g (mg L ) ) 0.4518 ± 0.002
nF 1.42 ± 0.08

us
2
R 0.968
R2adj 0.965

an
χ2 0.1147
Redlich–Peterson
KRP (L g-1) 1.8785 ± 0.04
M
aRP (mg L-1) 3.67 ± 0.01
g 0.31
d
2
R 0.967
2
R 0.960
te

adj

χ2 0.1846
*mean±standard error (n=3).
619
p

620
621
ce
Ac

29
Page 29 of 39
621 Table 5: Comparison of adsorption capacities (mg g–1) for the CV onto different adsorbents.
Equilibrium time Maximum adsorption capacity
Adsorbent Temperature (°C) Reference
(min) (mg g-1)
CPBP 120 30 32.78 [47]
Coir pith 60 25 24.75 [48]

t
Jackfruit leaf powder 120 20 43.39 [49]

ip
Peel cucumis sativa 90 27 34.24 [50]
Tomato plant root 15 30 94.33 [51]

cr
Turkish Coffee Residue 240 25 25.0 [52]
FPSP 60 25 85.99 This study

us
622
623

an
M
d
p te
ce
Ac

30
Page 30 of 39
Figure

Fig. 1

t
ip
cr
us
an
M
ed
pt
ce
Ac

Page 31 of 39
Figure

Fig. 2

10
(a)
pH final
pH initial
8 pHpzc= 6.851

t
6

ip
pHfinal

cr
2

us
0
0 2 4 6 8 10
pHinicial

(b)
an
10
M
5
pH zeta = 6.849

0
d
mV

-5
e
pt

-10

-15
ce

2 4 6 8 10 12

pH
Ac

Page 32 of 39
Figure

Fig. 3

a
b
Absorbance (a.u.)

t
ip
cr
us
4000 3500 3000
an
2500 2000 1500 1000
-1
Wavenumber (cm )
M
ed
pt
ce
Ac

Page 33 of 39
Figure

Fig. 4

A
(a)

t
ip
cr
us
(b)
B
an
M
ed
pt
ce
Ac

Page 34 of 39
Figure

Fig. 5

100
Wheigt loss (%)

80

t
ip
60

cr
40

us
20

100 200 300


an
400
o
Temperature ( C)
500 600 700
M
ed
pt
ce
Ac

Page 35 of 39
Figure

Fig. 6

100
(a)
90

80

t
70

ip
R(%)

60

cr
50

us
40

30
2 3 4 5
an
6 7 8 9 10
Initial pH of solution
M
ed

100 (b)

90
pt

80
ce
R(%)

70

60
Ac

50

40

30
2 4 6 8 10 12 14 16 18 20
Adsorbent dosage (g L-1)

Page 36 of 39
Figure

Fig. 7

11 2
10 (a) (b)
9
-1
8 10.0 mg L
-1
20.0 mg L
7 40.0 mgL
-1
qt (mg g )

qt (mg g )
-1

-1
1
5
4
3 Experimental points
2 Pseudo-first order
Pseudo-second order
1 Elovich

t
0 0

ip
0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (min) Time (mim)

cr
us
10
6 (c) (d)
9
5 8
an
7
4
6
qt (mg g )
qt (mg g )

-1
-1

3 5
4
2 Experimental points 3 Experimental points
ed

Pseudo-first order Pseudo-first order


2
1 Pseudo-second order Pseudo-second order
Elovich 1 Elovich
0 0
pt

0 50 100 150 200 250 300 0 50 100 150 200 250 300
Time (min) Time (min)
ce
Ac

Page 37 of 39
Figure

Fig. 8

50

40

t
30

ip
qe (mg g )
-1

cr
20
Experimental points
Langmuir

us
10 Freundlich
Redlich-Peterson

0
0 100 200 300
an
400 500
-1
600 700 800
Ce ( mg L )
M
ed
pt
ce
Ac

Page 38 of 39
Figure

Fig. 9

100
Desorption percentage (%)

90

80

t
70

ip
60

cr
50

us
40

30

20 an
0.05 0.10 0.20 0.50 1.00 1.50 2.00
Eluent Concentration (mol L-1)
M
ed
pt
ce
Ac

Page 39 of 39

You might also like