Journal Pre-proof Synthesized Bioadsorbent from Fish Scale for Chromium (III) Removal Firomsa Teshale, R. Karthikeyan, Omprakash Sahu
PII:
S0968-4328(19)30297-5
DOI:
https://doi.org/10.1016/j.micron.2019.102817
Reference:
JMIC 102817
To appear in:
Micron
Received Date:
14 September 2019
Revised Date:
24 December 2019
Accepted Date:
24 December 2019
Please cite this article as: Teshale F, Karthikeyan R, Sahu O, Synthesized Bioadsorbent from Fish Scale for Chromium (III) Removal, Micron (2019), doi: https://doi.org/10.1016/j.micron.2019.102817
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. 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. © 2019 Published by Elsevier.
Synthesized Bioadsorbent from Fish Scale for Chromium (III) Removal
Firomsa Teshale, R. Karthikeyan and Omprakash Sahu
Department of Chemical Engineering, KiOT, Wollo University Ethiopia Department of Chemical and BioEngineering, AAiT, Addis Ababa, Ethiopia
ro
Corresponding author Email:
[email protected]; Tel:+919752610957
ur na
lP
re
-p
Graphical-Abstract
Jo
of
Department of Chemical and Petroleum Engineering, UIE, Chandigarh University, Mohali, India
Highlights Achieving the permissible norm and avoiding the unpleasant odour
Bioadsorption preparation form waste fish scales
Characterisation of chemical synethised bioadsorbents before and after experiments
Deduction of hazardous metal ions from industrial effluent
Significant role of operating parameters on pollutant removal
ro
of
-p
Abstract:
re
Presence of heavy metal in industrial wastewater is hazardous to the surrounding environment. Biosorption of heavy metal is an effective technology for the treatment of industrial wastewater.
lP
This research work has been carried out on removal of chromium (III) metal ions by employing waste fish scales as bioadsorbent. A batch adsorption process was carried out with different
ur na
adsorbent dosage, solution pH and contact time. The results show the highest 99.7518 % chromium (III) metal ions at bioadsorbent dosage 0.8g, pH of the solution 5 and contact time 90 min, initial concentration 150mg/l chromium ion. The adsorption isotherms data fitted well with the Langmuir isotherm model with R2= 0.9998, qmax= 18.3486 mg/g, and RL= 0.00007325. As
Jo
well as pseudo-first and second kinetics model was also analyzed for the description of adsorption and found to be well fitted (R2=1) for adsorption kinetics. The surface properties activated fish scales and chromium loaded fish scale were investigated by scanning electron microscopy, X-ray spectroscopy, Fourier-transform infrared spectroscopy, and thermal analysis and agree with outcomes.
Keywords: Bioadsorbent; Effluent; Harmful; Metal ions; Toxic waste Introduction: Leather processing has emerged as an important economic activity in a developing country (Ethiopia) (Wakeford, et al., 2016). The leather goods market value forecast worldwide from 2018 to 2021 was nearly 239.78 to 271.21 billion USD (TSP-Report, 2018). It means increasing 10.47% per annum, which required a large amount of leather to fulfill the world demand.
of
Tannery industries are among those industries, which required a large quantity of water for
ro
processing and discharge 90% of water as effluent. Nearly 40-45 liters of fresh water were required to process one kilogram of raw leather from hide or skin (Sathish et al., 2016). The
-p
wastewater from tannery industries loaded with a high percentage of a color compound, various
re
organic and inorganic substances, toxic metallic compounds and different types of tanning materials (Chowdhury et al.,2015). Among all these pollutants, chromium that available in
lP
tannery effluent has significant health concerns due to high-ranking oxidizing properties (Magoling et al., 2017). Chromium is present in a trivalent and hexavalent form in an aqueous
ur na
phase and allowable limit is 0.1mg/L for surface water and 0.05mg/L for potable water, according to world health organization (WHO, 2011). Including to lathering industry other industries like electroplating, textile and dye manufacturing, metal plating, and battery production were also contributes chromium ion in the range of 0.5mg/L to 270 mg/L
Jo
(Kamsonlian and Shukla, 2013). Hence, it’s necessary to remove the chromium from tannery industries wastewater to make the environment safe and clean (Ghaedi and Mosallanejad, 2018). Various treatment methods have been suggested to reduce the level of chromium from
wastewater in literature like chemical precipitation (Gheju and Balcu, 2011), coagulation (Golbaz et al., 2014), electrocoagulation (Verma et al., 2013); solvent extraction (Vander
Hoogerstraete et al., 2013); electrolysis (Sadyrbaeva, 2016), membrane separation (Mohammed, K. and Sahu, 2015); ion–exchange (Kumar et al., 2017) and biological (Gutiérrez-Corona et al., 2016) etc. These methods are found to be less effective on chromium metal ion elimination, due to technical terms. As compared to all; the adsorption process has been suggested as effective in terms of economic and operation (Kobielska et al., 2018). Adsorption process has been used to different types of different industrial effluent like dye industries (Vikrant et al., 2018), fertilizer
of
(Melia et al., 2017); distillery (Ioannou et al., 2015); food (Qasim and Mane, 2013);
ro
pharmaceutical (Deegan et al., 2011) as well as heavy metal (Visa, 2016). The commercial activated carbons used in adsorption process make expensive treatment, for this reason, an
-p
alternative like bioadsorbent are focusing nowadays (Enniya et al., 2018). However different
re
types of bioadsorbent like Spirulina sp. (Rezaei,2016); Sugarcane bagasse (Ullah et al., 2013); Macrocystis pyrifera and Undaria pinnatifida (Cazón et al., 2012); Ulva sp. (Vijayaraghavan and
lP
Joshi, 2014); Orange peel (Lugo-Lugo et al., 2012); olive stone (Blazquez, et al., 2011); microalgae (Kim et al., 2011); vineyard pruning (Karaoğlu et al., 2010) has been reported for
ur na
chromium (III) ions removal. Bioadsorbents are biomass matter, which consists a range of functional groups like amino, carboxyl, hydroxyl and phosphate on the cell wall are magnetize the heavy metals on the process (Carolin et al., 2017; Adio et al., 2019). The biomass material can be effortlessly reprocessed and dispose of after regeneration. Nearly 143.8 million tonnes of
Jo
fish consumed per annum, which produced 50% as waste (Hollander et al., 2019). So afford has been made to develop low-cost bioadsorbent from waste fish scale for chromium (III) removal in order to meet permissible limits standards as well as to avoid unsafe disposal of waste fish scale.
The main aim of this research work is to remove chromium (III) metal ions from synthetic solution by waste fish scale as bioadsorbent. The experiment conducts in batch mode and process parameters were optimized by using surface response methodology. Bioadsorbent was characterized by scanning electron microscopy (SEM), X-ray spectroscopy (XRD), Fouriertransform infrared spectroscopy (FTIR) and thermal analysis (DTA-DTG-TG). Adsorption isotherms and kinetics models pseudo-first and second order were also studied.
of
Material and methods:
ro
Material:
Synethic solution: A stock synthetic solution was prepared with 3.77gram of chromium sulfate
-p
(Cr2 (SO4)3) in one liter of deionized water.
re
Bioadsorbent: The waste scale (fish) was arranged from Arbaminch Lake Ethiopia. Chemicals: A laboratory grade sulphuric acid, sodium hydroxide, calcium hydroxide and
lP
chromium sulphate purchase from chemical supplier Addis Ababa (Ethiopia). Methods:
ur na
Preparation of adsorbent: To prepare the bioadsorbent, arranged fish scale was washed and soaked for 24 hrs with normal water, and subsequently cleaned with ionized water to remove the impurities. Washed fish scales were kept in sundry for 48 hrs and open-air oven at 65oC for 24 hrs to make moisture free and crispy (Kongsri et al., 2013). The dried scales were converted into
Jo
a fine powder with help of mechanical grinder (Chen et al., 2010). The powder was sieved through octagon sieve and maintained the fraction size range of 125 to 200µm. The chemical activation was carried out with 10 gram of fish scale powder mixed with 150 ml 0.1M sulphuric acid solution (acid activation) for 2.5 hrs at 100 rpm rotary shaker in room temperature, similar procedures were followed with sodium hydroxide (base activation). The chemically treated
suspension was transferred to a vacuum filter to separate the liquid part (Iqba et al., 2011). Collected semi slurry was washed repeatedly using distilled water until a neutral pH solution was attained. Afterward, the sample was dried at room temperature for 2 hrs and on the oven at 80oC. The dried sample was kept in poly bags until used (Iqbal et al., 2011). The steps fallow to prepared bioadsorbent from waste fish scales is shown in Fig. 1. Experimental design:
of
A response surface methodology (RSM) was applied for experiment design on chromium
ro
removal in the batch process. In this research, adsorbent dosage, a range of 0.4g; 0.8g and 1g pH range of pH 3; 5 and 8 and contact time range 30; 90 and 120 mins, considered as a factor for
-p
three levels, which mentioned in Table 1. The response (Y) indicates the chromium removal in
re
percentage and factor coded in X1(bioadsorbent dosage); X2 (pH) and X3 (Contact time). The mathematical relationship between response and variables can be presented in an Eq. (1) and in
Y f ( x1 , x2 , x3 .........xn )
lP
the form of a second order polynomial (Witek-Krowiak et al., 2014): (1)
ur na
Where Y is the response of the system, and xi is the variables of action called factors. The independent variables are continuous and controllable by experiments with negligible errors has been assumed for an equation. The results of the Y (response) in an experiment by adsorption
Jo
were measured by design matrix listed in Table 2. The percentage of chromium reduction efficiency was calculated using the Eq. (2).
(2) Experiment-setup:
The adsorption process was carried out in batch mode to examine removal efficiency of activated fish scales (AFS) at room temperature. The experimental setup is presented in Fig.2. The initial pH of the solution between pH 3, 5 and 8 was adjusted by adding 0.1M hydrochloric acid (HCl) and sodium hydroxide (NaOH) solution. A known volume 100ml of synthetic chromium solution was taken in 250 ml beaker and measure amount of activated bioadsorbent 0.4g, 0.8g, and 1 g was added. The conical flask (sample and bioadsorbent) was kept in magnetic stirrer plate and
of
mixed at speed of 200 rpm for 30min; 90min and 150 min interval of time. After the adsorption
ro
process, separation of the sorbent and solutions was carried out by 90 mm diameter filter paper via vacuum filtration unit (Ding et al., 2012). The filtrate was analyzed for residual Cr (III) using
-p
atomic absorption spectrophotometer.
re
Analytical analysis:
The concentrations of chromium metal ions were determined using atomic absorption
lP
spectrophotometer. FT-IR spectroscopy was used to identify the functional groups in the fish scales biosorbent. FTIR spectra of the biosorbent before, after chemically modification and after
ur na
biosorption of Cr (III) loaded were recorded in the range 4000-400 cm-1 using spectrum 65FTIR (Perkin Elmer) KBr pellets. X-ray diffraction spectra were recorded, using a step size of 10o, step time of 1s and 2θ range of 10o to 70o. The samples powder, particles retained in the sieve with 150µm was used for characterization and analysis. Scanning Electron Microscope coupled
Jo
with Energy-dispersive X-ray spectrometer (SEM - EDX) (JSM – 7100F), was employed to determine the surface morphology of the fish scales and to determine its elemental composition (Nielsen et al., 2015). The TGA/DTA was performed on differential thermal analyzer of type (SII 6300 EXSTAR) to estimate the complete decomposition temperature of fish scales sample
by heating it from 25oC to 1100oC with 10oC/min of constant heating rate and under nitrogen flow as an inert gas. Result and discussion: Characterization of Fish Scales Fourier Transform Infrared spectrum Analysis The FTIR spectroscopy is an important analytical technique which identifies the vibration
of
characteristics of chemical functional groups in a fragment. The FTIR curves of raw fish
ro
scales, modified fish scale and chromium (III) loaded fish scales are shown in Fig. 3 (a), (b) and (c). The raw fish scales, the stretching broadband at 3365cm-1 corresponds to O-H groups.
-p
The region between 3000 and 2800cm-1 exhibited medium peak, indicates the C-H stretching
re
vibrations –CH3 and -CH2 functional groups (Ribeiro et al., 2015). Distinct strong peaks observed between 1715 and 1630 cm-1 characterize carbonyl groups strong stretching. From
lP
1300-1470 cm-1 there was the deformation stretching of CH, CH3, and CH2 functional groups as a well medium peak at 1030 cm-1, due to C–O group. Some other peaks of P-O were also
ur na
recorded at 750 cm-1 and 550 cm-1 due to PO43- groups, those are characteristic of the inorganic phase (Nadeem et al., 2008). Changes in intensity and shift in the position of the peaks could be observed in FT-IR spectrum after Cr (III) adsorption on modified fish scales (Fig. 3 (c)). The shifting of the peak at 1660 to 1655 cm-1indicates that the involvement of
Jo
C=O/C=C in the adsorption process. The strong peak of raw biosorbent at 1030cm-1 (Fig. 3(a) ranging from 1100 to 1020 cm-1 shifted to 1025cm1. This suggested the participation of C-O group in the binding of Cr (III). The deformation stretching peak between 1300 -1407 cm-1 (Fig. 3 (a)) was changed and PO43- medium peaks are changed to small peak (Fig. 3(c)). The result indicates that the hydroxyl and carbonyl groups are the main functional groups on the
surface of raw fish scales involved in chromium (III) uptake to the modified fish scale. The main functional groups present on the surface of synthesized fish scales were hydroxyl, carbonyl, and phosphate. These observations have good agreement with the FTIR data for Tilapia fish scales used to remove hazardous azo dye reported in the literature (Zhu et al., 2013). The FT-IR analysis, the possible mechanism of adsorption of chromium (III) ion on the fish scale may be due to physical adsorption, ion exchange, chemical reaction with surface
of
sites, and complexation with functional groups. This result also agree with the removal of
ro
chromium (III) by other bioadsorbent like rice husk and saw dust of Eucalyptus camaldulensis
et al., 2014) and pine bark (Arim et al., 2018) etc.
re
X-ray Diffraction (XRD) Analysis
-p
(Kanwal et al., 2012); olive-mill waste (Bautista-Toledo et al., 2014); graphene oxide (Yang
Generally X-ray diffraction phase/composition identification will distinguish the major, minor,
lP
and trace compounds present in a sample. X-ray diffraction was conducted for the surfaces of raw fish scales; synthesized scale and chromium loaded scale which shown in Fig. 4 (a); (b) and
ur na
(c). The XRD patterns with a 2θ scan from 20◦ to 70◦. The peaks at 2θ of 25.96o, 28.66o 31.66o, 32.05o, 39.71o, 44.22o, and 49.66o correspond to protein and calcium. The Bragg equation d(Å) = λ/2sinθ (λ = 1.54 Å), minimum values (d) of spacing were 0.3432, 0.3115, 0.2827, 0.2792, 0.2270, and 0.1836nm calculated. The broad peaks of the XRD pattern indicate that the
Jo
scale has low crystallinity which corresponds to an amorphous property for chromium loaded fish scale. It may be due to the significant interaction between chromium ions and calcium. This peaks are agreed with result peaked at 2θ of 25.9o, 31.9o, 39.8o, 47.2o, 49.3o, and 53.2o correspond to d spacing of 0.343, 0.281, 0.226, 0.193, 0.185, and 0.172 nm, reported by (Chakraborty et al., 2011) and peaked at 25.8o, 31.8o, 39.6o, 47.2o, 49.3o, and 53.1o corresponding
to d spacing of 0.345, 0.281, 0.227, 0.192, 0.184, and 0.172 nm (Torres et al., 2008). The chemical composition of raw fish scale is mention in Table 3. Scanning electron Micrograph: The surface structure of the activated fish scale and after adsorption scale was analyzed under a scanning electron microscope. The scanning electron micrographs of the activated fish scale (AFS) before and after the Cr (III) adsorption at 1000× magnification is shown in Fig. 5 (a) and
of
(b), respectively. It can be seen honeycomb-like structures or open pores are available on the
ro
surface of the activated fish scale. The open pores are actively participated during the process to entrap the metal ion on their surface. These honeycomb cages formed due to inorganic material
-p
calcium and oxygen and organic material protein present in fish scale (Stevens and Batlokwa,
re
2017). After adsorption fish scale layers have different from the native bioadsorbent. The surfaces are damaged, rough and irregular structure due to metal ion trapped on the pores. Hence
Thermal analysis:
lP
it confirms that physical adsorption occurred on the surface of AFS (Onwordi et al., 2019).
ur na
The thermal gravimetric analysis (TGA), differential thermal analysis (DTA) and differential thermal gravimetry (DTG) curves of activated fish scales (AFS) and chromium loaded fish scale (CLFS) absorbent in an oxidizing atmosphere is shown in Fig. 6 (a) and (b), respectively. The DTA trace shows dehydration and volatilization (removal of volatiles) of the sample up to 300oC
Jo
losing weight of 1.4% for activated fish scales and 0.04% for chromium loaded fish scale bioadsorbent. These losses occurred due to the moisture and less unstable material present on both bioadsorbent samples (Pati et al., 2010). The peak rate of weight loss 0.7mg/min was seen at temperature Tmax 780oC for activated fish scale and 105.2 µg at Tmax 716oC for chromium loaded fish scales. Oxidizing of solids was found exothermic, 2000J/g heat evolution for
activated fish scale at 780oC. The weight loss between 300oC to 796oC was 44.6% for AFS and 39.65% weight loss up-to 765oC temperature for chromium loaded fish scale, may be due to total loss of CO2 contained in bioadsorbent (Zhao et al., 2018). Oxidation seems to be completed at a temperature of 1108oC for AFS and 1110oC for CLFS due to decomposition and burning of organic components in both bioadsorbent almost bring to an end. At 1108oC the 54.51% weight of activated fish scales ash and 59.91% weight at 1110oC of chromium loaded fish scale as ash
of
found to remain. Thus chromium loaded fish scales contains more inorganic material as
ro
compared to the activated fish scales (Mondal et al., 2012). Statistical Analysis
-p
The process was analyzed for process parameters adsorbent dosage (X1), pH (X2), and contact
re
time (X3) on adsorption of chromium (III) (Y) by using design expert 7. The laboratory experiment was done at a constant initial concentration (150mg/L) of pollutant at 200rpm
lP
shaking speed and room temperature. The experiment results analysis was carried out by using analysis of variance of the quadratic regression model for analysis significant model
ur na
term (Srivastava et al., 2015). To determine whether or not the quadratic model is significant, it was required to perform an analysis of variance (ANOVA), mention Table 4. According to analysis model terms, probability factor (Prob. > F) values less than 0.05 are considered significant. In this work, the model terms X2, X3, X22, X32, and X1X2 are a significant model,
Jo
while X1 is least significant terms. The values greater than 0.1000 indicates the model terms X1X3 and X2X3 are not significant. The quality of the model developed was evaluated based on the correlation coefficient (R2) and standard deviation. The "Pred R2" of 0.4716 is reasonable agreement with “Adj R2" of 0.6665, which is less than 0.2, this might be due to signal to disturbance ratio. A ratio greater than 4 is desirable; here the ratio of 8.411 indicates
an adequate signal. Therefore, this model can be used to navigate or direct the design space. The regression coefficient and the corresponding 95% confidence interval (CI) high and low were mention in Table 5. From designed experimental data, the quadratic polynomial model equation for chromium (III)
ro
of
adsorption using a modified fish scale is mention in of finial coded factor on Eq. (3):
(3)
-p
The generated data were also investigated for normality of the residuals and relationship
re
between the actual and predicted values of response (Y), shown in Fig.7. The normal probability plot indicates that the residuals observed value is more close to normal
lP
distribution. These demonstrate that the experiment points in plots are fit to the straight line and the quadratic polynomial model satisfies the assumptions of analysis of variance (Mohan
ur na
et al., 2015).
Effect of functioning parameters:
The results of different functioning parameters are presented in Fig. 8 (a), (b) and (c). An initial pH of the solution is the most important parameter affecting metal ion adsorption. The
Jo
effect of pH on chromium absorption process was carried out at constant dosage 0.70g; time 90 min; initial Cr (III) concentration 150 mg/L and different range of pH 3 to pH 8. It was found that maximum chromium removal occurred at pH 5.8, which shown in Fig. 8 (a). The result also shows that with an increase or decrease or near to neutral pH of a solution, adsorption capacities were found to be low due to weak electrostatic force of attraction between the oppositely charged adsorbate and adsorbent (Subramaniam and Ponnusamy,
2015; Liang et al., 2018). At pH 5.8 to 6.3 rates of adsorption chromium was constant beyond this pH the efficiency of biosorbent was decreased due to increase in pH that causes a formation of soluble hydroxyl complexes and decreases adhering capacity of metal ions (Nazir et al., 2019). Almost 99.92% chromium (IV) removal was observed with Artocarpus heterophyllus peel as bioadsorbent at pH 2; 50 mg/L initial concentration, 0.3 g of biosorbent at 100 rpm, and 35 °C experiment temperature (Saranya et al., 2018).
of
The effects of adsorbent dosage on the chromium (III) removal at room temperature
ro
was investigated by varying the amount of sorbent from 0.4g to 1g in a stock solution having Cr (III) ion concentration 150mg/l at pH 5.5 and time 90 mins. The result presented in Fig.8
-p
(b). It can be observed that chromium percentage removal increases with increase in
re
bioadsorbent dosage from 0.4g to 0.8g. This might be due to the availability of more adsorbent sites and enhanced the free surface area (Aman et al., 2018). Addition of activated
lP
fish scale beyond 0.8g the percentage removal of chromium was more or less near to constant. This may be due to particulate interaction such as agglomeration or equilibrium between
ur na
chromium ions and the solid phase was reached (Huang et al., 2016; Anastopoulos et al., 2019). Hence the positive correlation between adsorbent dose and chromium removal can be related to increasing the adsorbent surface area and the availability of more adsorption sites (Wu et al., 2017). Similarly 5mg/L of fish scale bioadsorbent dose was found to be suitable to
Jo
remove 94.2% of chromium at initial concentration 50mg/L chromium; pH 5 and experiment time 5 hours (Prabu et al., 2012). The effect of contact time on chromium reduction was carried at different time interval
30 min to 150 min. An experiment conducted at constant dosage 0.70g and pH 5.50; initial concentration of 150 mg/L at room temperature (21oC), which shown in Fig.8(c). It observed that
removal efficiency sharply increase with an increase in time from 30 min to 80 min and then attains equilibrium at 110 min. This is probably due to the availability of larger surface area on activated fish scale surface, which prolonged contact between the adsorbent surface and adsorbate (metal ions) in solution (Ooi et al., 2017). The active sites of adsorbent were exhausted when equilibrium achieved the rate of uptake is controlled by the rate at which the adsorbate is transported from the exterior to the interior sites of the sorbent particles (Gogoi et al., 2018).
of
Lower adsorption efficiency rate afterward 110min may be due to challenges for Cr3+ ions to
ro
occupying the leftover vacant surface sites of adsorbents in bulk solution (Zare-Dorabei et al., 2016). Author reported 99% Cr (III) removal was noted at 14 hrs with eggshell and 30 min with
-p
marble powder as bioadsorbent respectively (Elabbas et al., 2016).
re
Interaction of process variables on chromium (III) Uptake
The interaction among the process variables was analyzed with surface response methodology
lP
and plotted three-dimensional graphs to optimize the chromium uptake. The 3D plotted and contour plot is presented in Fig.9 (a); (b) and (c).
ur na
Interaction of adsorbent dosage and solution pH The response (Y) of chromium (III) ions removal efficiency as a function of adsorbent dosage and solution pH, are represented as three dimensional and contour plot in Fig.9(a). The graph ascribes that with an increase in pH 3 of solution and adsorbent dosage 0.4mg/l, the
Jo
percentage of chromium removal efficiency increases to 96.43 % at invariable time 90.0 min. This efficiency increases up to 99.95 % chromium ions when sorbent dosage 0.8 mg/l and pH 5 increased at contact time. The increases in percentage removal of Cr3+ ion attributes to increase in an adsorbent dose that offered more surface area and completely consume the active site by an adsorbent at low pH (Daoud et al., 2015). Usually, as pH increases at the
lower level of adsorbent dosage and as adsorbent increases at the low level of pH gives a positive effect on the chromium removal efficiency (Liu et al., 2018). Interaction of adsorbent dosage and contact time The interaction of adsorbent dosage and contact time on chromium elimination efficiency were shown as three dimension and contour plots in Fig. 9(b). The results reveal that at a definite contact time 90 to 120 min, a performance of chromium removal increased and reached to peak
of
beyond 120 min of contact time performance decreases. However, increasing the dose beyond
ro
0.8g, there is gradual decreases removal efficiency of chromium ions by activated fish scale. The reason may be due sorbent particulate agglomeration, which rejects active site for adsorption
-p
surface area (Hu et al., 2016). Hence, maximum percentage removal of chromium (III) ion using
re
activated fish scale was achieved at a medium value of dosage and higher contact time 110120min. The plots indicate maximum adsorption 99.985% Cr3+ ion take place between 0.7 to 0.8
lP
g dosage and 110 to 120 min of contact time. Hence it’s confirmed that as contact time increase lower adsorbent dosage gives positive effects on the efficiency of chromium removal and
ur na
decreases in efficiency at a higher dose to lower contact time (Hashem et al., 2019). Interaction of solution pH and contact time The interaction of solution pH and contact time on chromium ions reduction is plotted on Fig. 9(c). The graph illustrated that contact time increases at the lower level of pH and pH value
Jo
increases at the low level of contact time have a significant impact on chromium (III) ions removal efficiency. It can be observed that as contact time increases up to 110 min, removal efficiency increased, additional of contact time beyond 110 min attains the stability and then decrease the performance of chromium removal. In the same way, when pH value increase from pH 3 to pH 5.8 chromium (III) removal efficiency increases, beyond the pH 5.8 value of
removal efficiency decreased. This might be due to dissociation of the biomass surface and solution chemistry of the heavy metals, which strongly affect with pH (Samuel et al., 2015). For this reason, when contact time increase at lower of pH and pH of the solution increases at lower contact time has beneficial effects on the removal efficiency of chromium (III) ions with activated fish scale (AFS) (Batool et al., 2019). Adsorption Isotherm
of
To examine the distribution of a solute between liquid and solid phase different mathematical
ro
relation such as standard Langmuir and Freundlich models has been suggested in the literature (Guyo et al., 2015). In this work, the isotherm studies were performed at initial concentration
-p
150mg/l, pH 5; adsorbent dosage 0.8g; contact time 2.5 hrs; shaking speed 200 rpm and
re
temperature ± 25oC. The biosorption equilibrium uptake capacity for each sample was
(4)
ur na
lP
calculated according to mass balance on the ions expressed in the Eq. (4).
Where V is the sample volume (L), C0 is the initial ion concentration (mg/L), Ce is the equilibrium or final ion concentration (mg/L), M is the biomass dry weight (g), and q e is the biomass biosorption equilibrium ions uptake capacity (mg/g).
Jo
Langmuir and Freundlich isotherms models were employed to describe the equilibrium between adsorbed ions on the sorbent (qe, q) and ions in the solution (Ce, q) in this research work.
Langmuir Isotherm Model: The Langmuir isotherm is valid for monolayer adsorption onto a surface containing a finite number of identical sites. This describes quantitatively the formation of a monolayer adsorbate on the outer surface of the adsorbent, and after that no further adsorption takes place. Based upon these assumptions, Langmuir derived in linear from represented, mention
of
in Eq.(5):
ro
(5)
Where, Ce, qe, qmax and KL representing the equilibrium concentration of chromium (III),
-p
adsorption capacity of activated fish scales at equilibrium, the maximum capacity of
re
bioadsorbent and Langmuir constant respectively. The qmax and RL in the Langmuir isotherm can be determined by plotting (Ce/qe) versus (Ce), which shown in Fig.10 (a). The essential
lP
characteristics of Langmuir isotherm can be expressed by the dimensionless constant called
ur na
equilibrium parameter RL, that mention in Eq. (6):
(6)
Where Co is the initial concentration of metal ion (mg/l), RL is a dimensionless equilibrium parameter. The characteristics of the RL value indicate the nature of biosorption: - unfavorable
Jo
(RL> 1), linear (RL = 1), favorable (0 < RL< 1) and irreversible (RL = 0) (Reddi et al., 2017). Freundlich isotherm model
The Freundlich isotherm model is describing the adsorption of solutes from an aqueous phase to a solid surface. This isotherm model assumes that the uptake of metal ions occurs on a heterogeneous surface by multilayer adsorption and that the amount of adsorbate adsorbed increases infinitely with an increase in concentration (Karimifard et al., 2016). Freundlich
adsorption isotherm is the relationship between the amounts adsorbed per unit mass of adsorbent, qe, and the concentration of Cr (III) at equilibrium, Ce. The linear form of the logarithmic equation for Freundlich isotherm is written in Eq. (7)
(7)
of
Where Kf and n are the Freundlich constants, Kf and n are the indicators of the adsorption
ro
capacity and adsorption intensity respectively. In this case, the plot of log C e vs log qe was employed to generate the intercept value of Kf and the slope of = 1/ n. Freundlich biosorption
-p
models for the removal of chromium (III) on the activated surface is shown in Fig. 10 (b). The calculated qmax, KL, and correlation coefficients [R2] for the Langmuir equation and RL
re
values and Kf, 1/n, and [R2] for the Freundlich equation at constant temperatures are mentions
lP
in Table 6. The regression coefficient (R2) for Langmuir and Freundlich isotherms was 0.9998, and 0.9599, respectively. Therefore, the result of correlation coefficient R2
ur na
comparison of Langmuir and Freundlich isotherm models, the Langmuir adsorption isotherm is the fit model for chromium (III) ion adsorption onto the modified fish scale sorbent. The Langmuir isotherm (R2 = 0.9998) fits the experimental data very well, this describes that homogeneous distribution of active sites onto AFS surface. The adsorbent, Langmuir and
Jo
Freundlich adsorption isotherms were plotted and obtained correlation coefficients. It can be concluded that adsorption of ions studied on prepared adsorbent gives correlation coefficient higher than 0.9998 follows by Langmuir isotherm (Xu et al., 2015).
Adsorption kinetics model The rate kinetics of Cr (III) adsorption by the modified fish scales adsorbent was analyzed using the pseudo-first-order and second-order rate kinetics model. The kinetics of chromium
(III) adsorption was studied from the time versus percentage of removal curves (mg/g), shown in Fig. 11 (a). The relationship time and amount of chromium adsorbed per adsorbent (mg/g) which is used to analyze adsorption kinetics model at initial concentration 150mg/l, pH 5, adsorbent dosage 0.8g, contact time varies from 30 min to 130min. Pseudo-first-order Kinetic Model In order to investigate the mechanism of adsorption, the pseudo-first-order equations were
of
used to test experimental data of initial concentrations. The pseudo-first-order kinetic
(8)
re
-p
ro
equation is presented in Eq (8):
Where, qe and qt are the adsorption capacities at equilibrium and time t respectively (mg/g) and
ur na
into Eq (9):
lP
k1 is the rate constant of pseudo-first-order adsorption (mn-1). The equation can be rearranged
(9)
The values of log (qe - qt) were linearly correlated with t. The plot of log (qe - qt) vs. t should
Jo
give a linear relationship. The first-order rate constant k1 (1/min) and the equilibrium capacity qe can be obtained from the slope and intercept, respectively, which shown in Fig. 11 (b). The observed rate constant for the removal of Cr (III) ions on AFS for Pseudo-first-order kinetic model was described in Table 7. Pseudo-second-order Kinetic Model The pseudo-second-order adsorption kinetic is expressed in Eq (10):
(10) For the boundary conditions t=0 to t=t and qt= 0 and qt=qt, the integrated form of the above equation will be written in Eq. (11).
ro
The Eq.(11) has been rearranged to obtain a linear form of Eq (12):
of
(11)
(12)
-p
Where qe and qt are the sorption capacity at equilibrium and time t (mg/g), respectively, k 2 is the rate constant of the pseudo-second-order sorption (g/mg min). The k2 and qe are found
re
from the intercept and slope of t/qt versus t linear plot such that qe = 1/slope and k2 =
lP
slope2/intercept. The second order kinetic plot is shown in Fig. 11 (c). The comparison of experimental sorption capacities (qexp) and the predicted values (qcal, K1, K2,
ur na
and R2) from the pseudo first order and second order are given in Table 7. According to results pseudo-second-order kinetic equations having high R2 (R2= 1) values and sorption capacity qe (mg/g) as well as calculated and experimental value is near to same value (Asfaram et al., 2015). Hence second-order kinetics models were considered as the fit equations for bioadsorption of
Jo
chromium (III) ion using activated fish scales. Comparatively, study: Present research work for chromium (III) ions removal was compared with other research findings, which is mention in Table 8. Form the all the result it has been concluding that activated fish scale was found to be superior to eliminate the chromium (III) ion from tannery wastewater. In this research work pH is near to natural (pH=5); other research removed at the highly acidic solution (pH 2 to pH 4). The treatment time is 90 mins,
but others experimental time is almost more than 120 min. The efficiency 99.75% was also highest among all the finding till time. Conclusion Removal of chromium (III) form of tannery wastewater solutions was possible with waste fish scale as bioadsorbent. The result indicates that activated fish scales have great potential to removal Cr(III) from tannery wastewater. An experimental output confirms that using
of
activated fish scales 99.7518% chromium metal ion can be achieved at adsorbent dosage 0.8g,
ro
pH 5, and contact time 90 min and initial concentration 150mg/l. The adsorption process of Cr(III) depends on pH and contact time and justified the removal efficiency increases till
-p
equilibrium. The FTIR analysis describes that the adsorbent fish scales contain carboxyl,
re
amine, and hydroxyl functional groups. Based on the FTIR spectra, these functional groups are participating on adsorption of chromium on activated fish scales surface. The X-ray
lP
diffraction analysis indicates that fish scales have properties of an amorphous, these results agree with scanning electron micrograph and thermal analysis. The adsorption kinetics studies
ur na
indicated at equilibrium attained at 130min of contact between the activated fish scale and chromium solution and followed by pseudo-second-order. The Langmuir and Freundlich isotherm were studied and Langmuir isotherms showed a better fit to the process R2= 0.9998. In future, it can study real tannery wastewater and regeneration of bioadsorbent. With the
Jo
above result it can be concluding that employing waste fish scale for chromium removal is best alternative to minimise the pollution from surrounding. Acknowledgment: Authors acknowledge to Environment chair, School of Chemical and Bio Engineering, Addis Ababa Institute of Technology, Addis Ababa Ethiopia for providing lab facilities.
References: Adio, S.O., Asif, M., Mohammed, A.R.I., Baig, N., Al-Arfaj, A.A. and Saleh, T.A., 2019. Poly (amidoxime) modified magnetic activated carbon for chromium and thallium adsorption: statistical analysis and regeneration. Process Safety and Environmental Protection, 121, pp.254-262. Aman, A., Ahmed, D., Asad, N., Masih, R. and Abd ur Rahman, H.M., 2018. Rose biomass as a
of
potential biosorbent to remove chromium, mercury and zinc from contaminated waters.
ro
International Journal of Environmental Studies, 75(5), pp.774-787.
Anastopoulos, I., Robalds, A., Tran, H.N., Mitrogiannis, D., Giannakoudakis, D.A., Hosseini-
-p
Bandegharaei, A. and Dotto, G.L., 2019. Removal of heavy metals by leaves-derived
re
biosorbents. Environmental Chemistry Letters, 17(2), pp.755-766. Arim, A.L., Neves, K., Quina, M.J. and Gando-Ferreira, L.M., 2018. Experimental and
lP
mathematical modelling of Cr (III) sorption in fixed-bed column using modified pine bark. Journal of cleaner production, 183, pp.272-281.
ur na
Asfaram, A., Ghaedi, M., Goudarzi, A. and Rajabi, M., 2015. Response surface methodology approach for optimization of simultaneous dye and metal ion ultrasound-assisted adsorption onto Mn doped Fe 3 O 4-NPs loaded on AC: kinetic and isothermal studies. Dalton Transactions, 44(33), pp.14707-14723.
Jo
Batool, S., Idrees, M., Al-Wabel, M.I., Ahmad, M., Hina, K., Ullah, H., Cui, L. and Hussain, Q., 2019. Sorption of Cr (III) from aqueous media via naturally functionalized microporous biochar: mechanistic study. Microchemical Journal, 144, pp.242-253.
Bautista-Toledo, M.I., Rivera-Utrilla, J., Ocampo-Pérez, R., Carrasco-Marín, F. and SánchezPolo, M., 2014. Cooperative adsorption of bisphenol-A and chromium (III) ions from water on activated carbons prepared from olive-mill waste. Carbon, 73, pp.338-350. Blázquez, G., Calero, M., Hernáinz, F., Tenorio, G. and Martín‐ Lara, M.A., 2011. Batch and continuous packed column studies of chromium (III) biosorption by olive stone. Environmental Progress & Sustainable Energy, 30(4), pp.576-585.
of
Cazón, J.P.H., Benítez, L., Donati, E. and Viera, M., 2012. Biosorption of chromium (III) by two
ro
brown algae Macrocystis pyrifera and Undaria pinnatifida: Equilibrium and kinetic study. Engineering in Life Sciences, 12(1), pp.95-103.
-p
Chakraborty, R., Bepari, S. and Banerjee, A., 2011. Application of calcined waste fish (Labeo
technology, 102(3), pp.3610-3618.
re
rohita) scale as low-cost heterogeneous catalyst for biodiesel synthesis. Bioresource
lP
Chen, B., Wang, Y. and Hu, D., 2010. Biosorption and biodegradation of polycyclic aromatic hydrocarbons in aqueous solutions by a consortium of white-rot fungi. Journal of
ur na
hazardous materials, 179(1-3), pp.845-851.
Chowdhury, M., Mostafa, M.G., Biswas, T.K., Mandal, A. and Saha, A.K., 2015. Characterization of the effluents from leather processing industries. Environmental Processes, 2(1), pp.173-187.
Jo
Daoud, W., Ebadi, T. and Fahimifar, A., 2015. Optimization of hexavalent chromium removal from aqueous solution using acid-modified granular activated carbon as adsorbent through response surface methodology. Korean Journal of Chemical Engineering, 32(6), pp.11191128.
Deegan, A.M., Shaik, B., Nolan, K., Urell, K., Oelgemöller, M., Tobin, J. and Morrissey, A., 2011. Treatment options for wastewater effluents from pharmaceutical companies. International Journal of Environmental Science & Technology, 8(3), pp.649-666. Ding, Y., Jing, D., Gong, H., Zhou, L. and Yang, X., 2012. Biosorption of aquatic cadmium (II) by unmodified rice straw. Bioresource Technology, 114, pp.20-25. Elabbas, S., Mandi, L., Berrekhis, F., Pons, M.N., Leclerc, J.P. and Ouazzani, N., 2016. Removal
of
of Cr (III) from chrome tanning wastewater by adsorption using two natural carbonaceous
ro
materials: eggshell and powdered marble. Journal of environmental management, 166, pp.589-595.
-p
Eladlani, N., Ouahrouch, A., Rhazi, M. and Taourirte, M., 2018. Recovery of chromium (III)
re
from tannery wastewater by nanoparticles and whiskers of chitosan. Journal of Polymers and the Environment, 26(1), pp.152-157.
lP
Enniya, I., Rghioui, L. and Jourani, A., 2018. Adsorption of hexavalent chromium in aqueous solution on activated carbon prepared from apple peels. Sustainable Chemistry and
ur na
Pharmacy, 7, pp.9-16.
Ferraz, A.I., Amorim, C., Tavares, T. and Teixeira, J.A., 2015. Chromium (III) biosorption onto spent grains residual from brewing industry: equilibrium, kinetics and column studies. International Journal of Environmental Science and Technology, 12(5), pp.1591-
Jo
1602.
Ghaedi, M. and Mosallanejad, N.A.R.G.E.S., 2018. Removal of heavy metal ions from polluted waters by using of low cost adsorbents. Journal of Chemical Health Risks, 3(1) pp 21-30.
Gheju, M. and Balcu, I., 2011. Removal of chromium from Cr (VI) polluted wastewaters by reduction with scrap iron and subsequent precipitation of resulted cations. Journal of hazardous materials, 196, pp.131-138. Gogoi, S., Chakraborty, S. and Saikia, M.D., 2018. Surface modified pineapple crown leaf for adsorption of Cr (VI) and Cr (III) ions from aqueous solution. Journal of Environmental Chemical Engineering, 6(2), pp.2492-2501.
of
phenol,
chromium,
and
cyanide
from
aqueous
solution
by
ro
removal
of
Golbaz, S., Jafari, A.J., Rafiee, M. and Kalantary, R.R., 2014. Separate and simultaneous
coagulation/precipitation: Mechanisms and theory. Chemical Engineering Journal, 253,
-p
pp.251-257.
re
Gonçalves, A.C., Nacke, H., Schwantes, D., Campagnolo, M.A., Miola, A.J., Tarley, C.R.T., Dragunski, D.C. and Suquila, F.A.C., 2017. Adsorption mechanism of chromium (III)
lP
using biosorbents of Jatropha curcas L. Environmental Science and Pollution Research, 24(27), pp.21778-21790.
ur na
Gutiérrez-Corona, J.F., Romo-Rodríguez, P., Santos-Escobar, F., Espino-Saldaña, A.E. and Hernández-Escoto, H., 2016. Microbial interactions with chromium: basic biological processes and applications in environmental biotechnology. World Journal of Microbiology and Biotechnology, 32(12), p.191.
Jo
Guyo, U., Makawa, T., Moyo, M., Nharingo, T., Nyamunda, B.C. and Mugadza, T., 2015. Application of response surface methodology for Cd (II) adsorption on maize tasselmagnetite nanohybrid adsorbent. Journal of Environmental Chemical Engineering, 3(4), pp.2472-2483.
Hashem, M.A., Momen, M.A., Hasan, M., Nur-A-Tomal, M.S. and Sheikh, M.H., 2019. Chromium removal from tannery wastewater using Syzygium cumini bark adsorbent. International journal of environmental science and technology, 16(3), pp.1395-1404. Hollander, A., De Jonge, R., Biesbroek, S., Hoekstra, J. and Zijp, M.C., 2019. Exploring solutions for healthy, safe, and sustainable fatty acids (EPA and DHA) consumption in The Netherlands. Sustainability Science, 14(2), pp.303-313.
of
Hu, X., Wang, H. and Liu, Y., 2016. Statistical analysis of main and interaction effects on Cu (II)
ro
and Cr (VI) decontamination by nitrogen–doped magnetic graphene oxide. Scientific reports, 6, p.34378.
-p
Ioannou, L.A., Puma, G.L. and Fatta-Kassinos, D., 2015. Treatment of winery wastewater by
materials, 286, pp.343-368.
re
physicochemical, biological and advanced processes: a review. Journal of hazardous
lP
Iqbal, J., Wattoo, F.H., Wattoo, M.H.S., Malik, R., Tirmizi, S.A., Imran, M. and Ghangro, A.B., 2011. Adsorption of acid yellow dye on flakes of chitosan prepared from fishery wastes.
ur na
Arabian Journal of Chemistry, 4(4), pp.389-395. Kamsonlian, S. and Shukla, B., 2013. Optimization of Process Parameters using Response Surface Methodology (RSM): Removal of Cr (VI) from Aqueous Solution by Wood Apple Shell Activated Carbon (WASAC). Research Journal of Chemical Sciences 3(7), 31-37.
Jo
Kanwal, F., Rehman, R., Mahmud, T., Anwar, J. and Ilyas, R., 2012. Isothermal and thermodynamical modeling of chromium (III) adsorption by composites of polyaniline with rice husk and saw dust. Journal of the Chilean Chemical Society, 57(1), pp.10581063.
Karaoğlu, M.H., Zor, Ş. and Uğurlu, M., 2010. Biosorption of Cr (III) from solutions using vineyard pruning waste. Chemical Engineering Journal, 159(1-3), pp.98-106. Karimifard, S. and Moghaddam, M.R.A., 2016. Enhancing the adsorption performance of carbon nanotubes with a multistep functionalization method: optimization of Reactive Blue 19 removal through response surface methodology. Process Safety and Environmental Protection, 99, pp.20-29.
of
Kim, E.J., Park, S., Hong, H.J., Choi, Y.E. and Yang, J.W., 2011. Biosorption of chromium (Cr
ro
(III)/Cr (VI)) on the residual microalga Nannochloris oculata after lipid extraction for biodiesel production. Bioresource technology, 102(24), pp.11155-11160.
-p
Kobielska, P.A., Howarth, A.J., Farha, O.K. and Nayak, S., 2018. Metal–organic frameworks for
re
heavy metal removal from water. Coordination Chemistry Reviews, 358, pp.92-107. Kongsri, S., Janpradit, K., Buapa, K., Techawongstien, S. and Chanthai, S., 2013.
lP
Nanocrystalline hydroxyapatite from fish scale waste: Preparation, characterization and application for selenium adsorption in aqueous solution. Chemical engineering journal,
ur na
215, pp.522-532.
Kumar, P., Pournara, A., Kim, K.H., Bansal, V., Rapti, S. and Manos, M.J., 2017. Metal-organic frameworks: challenges and opportunities for ion-exchange/sorption applications. Progress in Materials Science, 86, pp.25-74.
Jo
Li, C., Huang, B., Li, C., Chen, X. and Huang, Y., 2016. In situ formation of nanoscale zerovalue iron on fish-scale-based porous carbon for Cr (VI) adsorption. Water Science and Technology, 73(9), pp.2237-2243.
Liang, X., Fan, X., Li, R., Li, S., Shen, S. and Hu, D., 2018. Efficient removal of Cr (VI) from water by quaternized chitin/branched polyethylenimine biosorbent with hierarchical pore structure. Bioresource technology, 250, pp.178-184. Liu, Q., Liu, Q., Liu, B., Hu, T., Liu, W. and Yao, J., 2018. Green synthesis of tanninhexamethylendiamine based adsorbents for efficient removal of Cr (VI). Journal of hazardous materials, 352, pp.27-35.
of
Lugo-Lugo, V., Barrera-Díaz, C., Ureña-Núñez, F., Bilyeu, B. and Linares-Hernández, I., 2012.
ro
Biosorption of Cr (III) and Fe (III) in single and binary systems onto pretreated orange peel. Journal of Environmental Management, 112, pp.120-127.
-p
Magoling, B.J.A. and Macalalad, A.A., 2017. Optimization and Response Surface Modelling of
re
Activated Carbon Production from Mahogany Fruit Husk for Removal of Chromium (VI) from Aqueous Solution. BioResources, 12(2), pp.3001-3016.
lP
Magsi, S.K., Kandhar, I.A., Brohi, R.O.Z. and Channa, A., 2019, July. Removal of metals from water using fish scales as a bio adsorbent. In AIP Conference Proceedings (Vol. 2119,
ur na
No. 1, p. 020023). AIP Publishing.
Melia, P.M., Cundy, A.B., Sohi, S.P., Hooda, P.S. and Busquets, R., 2017. Trends in the recovery of phosphorus in bioavailable forms from wastewater. Chemosphere, 186, pp.381-395.
Jo
Mohammed, K. and Sahu, O., 2015. Bioadsorption and membrane technology for reduction and recovery of chromium from tannery industry wastewater. Environmental Technology & Innovation, 4, pp.150-158.
Mohan, S., Singh, Y., Verma, D.K. and Hasan, S.H., 2015. Synthesis of CuO nanoparticles through green route using Citrus limon juice and its application as nanosorbent for Cr (VI)
remediation: Process optimization with RSM and ANN-GA based model. Process Safety and Environmental Protection, 96, pp.156-166. Mondal, S., Mondal, B., Dey, A. and Mukhopadhyay, S.S., 2012. Studies on processing and characterization of hydroxyapatite biomaterials from different bio wastes. Journal of Minerals and Materials Characterization and Engineering, 11(01), p.55. Nadeem, R., Ansari, T.M. and Khalid, A.M., 2008. Fourier Transform Infrared Spectroscopic
of
characterization and optimization of Pb (II) biosorption by fish (Labeo rohita) scales.
ro
Journal of Hazardous Materials, 156(1-3), pp.64-73.
Nazir, M.S., Tahir, Z., Akhtar, M.N. and Abdullah, M.A., 2019. Biosorbents and Composite
-p
Cation Exchanger for the Treatment of Heavy Metals. In Applications of Ion Exchange
re
Materials in the Environment (pp. 135-159). Springer, Cham.
Nielsen, L., Zhang, P. and Bandosz, T.J., 2015. Adsorption of carbamazepine on sludge/fish
Journal, 267, pp.170-181.
lP
waste derived adsorbents: Effect of surface chemistry and texture. Chemical Engineering
ur na
Onwordi, C.T., Uche, C.C., Ameh, A.E. and Petrik, L.F., 2019. Comparative study of the adsorption capacity of lead (II) ions onto bean husk and fish scale from aqueous solution. Journal of Water Reuse and Desalination. Ooi, J., Lee, L.Y., Hiew, B.Y.Z., Thangalazhy-Gopakumar, S., Lim, S.S. and Gan, S., 2017.
Jo
Assessment of fish scales waste as a low cost and eco-friendly adsorbent for removal of an azo dye: Equilibrium, kinetic and thermodynamic studies. Bioresource technology, 245, pp.656-664.
Pati, F., Adhikari, B. and Dhara, S., 2010. Isolation and characterization of fish scale collagen of higher thermal stability. Bioresource technology, 101(10), pp.3737-3742.
Prabu, K., Shankarlal, S. and Natarajan, E., 2012. A biosorption of heavy metal ions from aqueous solutions using fish scale (Catla catla). World Journal of Fish and Marine Sciences, 4(1), pp.73-77. Qasim, W. and Mane, A.V., 2013. Characterization and treatment of selected food industrial effluents by coagulation and adsorption techniques. Water Resources and Industry, 4, pp.112.
of
Rezaei, H., 2016. Biosorption of chromium by using Spirulina sp. Arabian Journal of Chemistry,
ro
9(6), pp.846-853.
Ribeiro, C., Scheufele, F.B., Espinoza-Quinones, F.R., Módenes, A.N., da Silva, M.G.C., Vieira,
-p
M.G.A. and Borba, C.E., 2015. Characterization of Oreochromis niloticus fish scales and
re
assessment of their potential on the adsorption of reactive blue 5G dye. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 482, pp.693-701.
lP
Sadyrbaeva, T.Z., 2016. Removal of chromium (VI) from aqueous solutions using a novel hybrid liquid membrane—electrodialysis process. Chemical Engineering and Processing: Process
ur na
Intensification, 99, pp.183-191.
Samuel, M.S., Abigail, M.E.A. and Ramalingam, C., 2015. Biosorption of Cr (VI) by Ceratocystis paradoxa MSR2 using isotherm modelling, kinetic study and optimization of batch parameters using response surface methodology. PloS one, 10(3), p.e0118999.
Jo
Saranya, N., Ajmani, A., Sivasubramanian, V. and Selvaraju, N., 2018. Hexavalent chromium removal from simulated and real effluents using Artocarpus heterophyllus peel biosorbent-batch and continuous studies. Journal of Molecular Liquids, 265, pp.779-790.
Sathish, M., Madhan, B., Sreeram, K.J., Rao, J.R. and Nair, B.U., 2016. Alternative carrier medium for sustainable leather manufacturing–a review and perspective. Journal of cleaner production, 112, pp.49-58. Silva, E.A., Vaz, L.G.L., Veit, M.T., Fagundes-Klen, M.R., Cossich, E.S., Tavares, C.R.G., Cardozo-Filho, L. and Guirardello, R., 2010. Biosorption of chromium (III) and copper (II) ions onto marine alga Sargassum sp. in a fixed-bed column. Adsorption Science &
of
Technology, 28(5), pp.449-464.
ro
Srivastava, V., Sharma, Y.C. and Sillanpää, M., 2015. Response surface methodological approach for the optimization of adsorption process in the removal of Cr (VI) ions by Cu2
-p
(OH) 2CO3 nanoparticles. Applied Surface Science, 326, pp.257-270.
re
Stevens, M.G. and Batlokwa, B.S., 2017. Environmentally friendly and cheap removal of lead (II) and zinc (II) from wastewater with fish scales waste remains. International Journal of
lP
Chemistry, 9(4), p.22.
Subramaniam, R. and Ponnusamy, S.K., 2015. Novel adsorbent from agricultural waste (cashew
ur na
NUT shell) for methylene blue dye removal: optimization by response surface methodology. Water Resources and Industry, 11, pp.64-70. Sulaymon, A.H., Mohammed, A.A. and Al-Musawi, T.J., 2014. Comparative study of removal of cadmium (II) and chromium (III) ions from aqueous solution using low-cost biosorbent.
Jo
International Journal of Chemical Reactor Engineering, 12(1), pp.477-486.
The
Statistics
Portal-Report-2018:https://www.statista.com/statistics/861562/leather-goods-
market-value-worldwide/
Torres, F.G., Troncoso, O.P., Nakamatsu, J., Grande, C.J. and Gomez, C.M., 2008. Characterization of the nanocomposite laminate structure occurring in fish scales from Arapaima gigas. Materials Science and Engineering: C, 28(8), pp.1276-1283. Ullah, I., Nadeem, R., Iqbal, M. and Manzoor, Q., 2013. Biosorption of chromium onto native and immobilized sugarcane bagasse waste biomass. Ecological Engineering, 60, pp.99107.
of
Vander Hoogerstraete, T., Wellens, S., Verachtert, K. and Binnemans, K., 2013. Removal of
ro
transition metals from rare earths by solvent extraction with an undiluted phosphonium ionic liquid: separations relevant to rare-earth magnet recycling. Green Chemistry, 15(4),
-p
pp.919-927.
re
Verma, S.K., Khandegar, V. and Saroha, A.K., 2013. Removal of chromium from electroplating industry effluent using electrocoagulation. Journal of Hazardous, Toxic, and Radioactive
lP
Waste, 17(2), pp.146-152.
Vijayaraghavan, K. and Joshi, U.M., 2014. Application of Ulva sp. biomass for single and binary
ur na
biosorption of chromium (III) and manganese (II) ions: equilibrium modeling. Environmental Progress & Sustainable Energy, 33(1), pp.147-153. Vikrant, K., Giri, B.S., Raza, N., Roy, K., Kim, K.H., Rai, B.N. and Singh, R.S., 2018. Recent advancements in bioremediation of dye: current status and challenges. Bioresource
Jo
technology.
Visa, M., 2016. Synthesis and characterization of new zeolite materials obtained from fly ash for heavy metals removal in advanced wastewater treatment. Powder Technology, 294, pp.338-347.
Wakeford, J.J., Gebreeyesus, M., Ginbo, T., Yimer, K., Manzambi, O., Black, M., Mulugetta, Y. and Okerekee, C., 2016. An Assessment of Ethiopia’s Innovation Systems in Relation to Green Industrialisation. Witek-Krowiak, A., Chojnacka, K., Podstawczyk, D., Dawiec, A. and Pokomeda, K., 2014. Application of response surface methodology and artificial neural network methods in modelling and optimization of biosorption process. Bioresource technology, 160, pp.150-
of
160.
Health Organization Press, Geneva, Switzerland.
ro
World Health Organization. (2011). Guidelines for Drinking-Water Quality (4th ed.), World
-p
Wu, Y., Ming, Z., Yang, S., Fan, Y., Fang, P., Sha, H. and Cha, L., 2017. Adsorption of
re
hexavalent chromium onto Bamboo Charcoal grafted by Cu2+-N-aminopropylsilane complexes: Optimization, kinetic, and isotherm studies. Journal of industrial and
lP
engineering chemistry, 46, pp.222-233.
Xu, C., Wang, J., Yang, T., Chen, X., Liu, X. and Ding, X., 2015. Adsorption of uranium by
ur na
amidoximated chitosan-grafted polyacrylonitrile, using response surface methodology. Carbohydrate polymers, 121, pp.79-85. Yang, S., Li, L., Pei, Z., Li, C., Lv, J., Xie, J., Wen, B. and Zhang, S., 2014. Adsorption kinetics, isotherms and thermodynamics of Cr (III) on graphene oxide. Colloids and Surfaces A:
Jo
Physicochemical and Engineering Aspects, 457, pp.100-106.
Zare-Dorabei, R., Ferdowsi, S.M., Barzin, A. and Tadjarodi, A., 2016. Highly efficient simultaneous ultrasonic-assisted adsorption of Pb (II), Cd (II), Ni (II) and Cu (II) ions from aqueous solutions by graphene oxide modified with 2, 2′-dipyridylamine: central composite design optimization. Ultrasonics sonochemistry, 32, pp.265-276.
Zhao, R., Deng, S., Wang, S., Zhao, L., Zhang, Y., Liu, B., Li, H. and Yu, Z., 2018. Thermodynamic research of adsorbent materials on energy efficiency of vacuum-pressure swing adsorption cycle for CO2 capture. Applied Thermal Engineering, 128, pp.818-829. Zhu, K., Gong, X., He, D., Li, B., Ji, D., Li, P., Peng, Z. and Luo, Y., 2013. Adsorption of Ponceau 4R from aqueous solutions using alkali boiled Tilapia fish scales. RSC Advances,
Jo
ur na
lP
re
-p
ro
of
3(47), pp.25221-25230.
of ro -p re lP ur na Jo Fig.1:
Preparation
of
bioadsorbent
from
waste
fish
scales
of ro -p
Jo
ur na
lP
re
Fig.2: Experimental arrangement of adsorption process for chromium removal
of ro -p re lP ur na Jo Fig.3: Fourier Transform Infrared spectrum Analysis of (a) raw fish scales; (b) chemical modified scales and (c) chromium loaded scales
of ro -p re lP ur na Jo Fig.4: X-ray diffraction of (a) raw fish scales; (b) activated scales and (c) chromium loaded scale
of ro -p re lP ur na
Jo
Fig.5: Scanning electron micrograph analysis of fish scales (a) before adsorption process and (b) after adsorption process
re
lP
ur na
Jo -p
ro
of
of ro -p re lP ur na Jo
Fig.6: DTA-DTG-TG curve of (a) activated fish scale and (b) chromium loaded fish scale
of ro -p
Jo
ur na
lP
re
Fig.7: actual and predicted values of chromium removal
of ro -p re lP ur na Jo Fig.8: Effect of different functioning parameters on (a) initial pH; (b) dosage of bioadsorbent and (c) contact time
of ro -p re lP ur na
Jo
Fig. 9(a): Interaction three dimension plot and contour plot of adsorbent dosage and solution pH
of ro -p re lP ur na
Jo
Fig. 9(b): Interaction three dimension plot and contour plot of adsorbent dosage and contact time
of ro -p re lP ur na
Jo
Fig. 9(c): Interaction three dimension plot and contour plot of solution pH and contact time
of ro -p re lP ur na Jo
Fig. 10: Adsorption isotherms of chromium (a) Langmuir and (b) Freundlich models
of ro -p re lP ur na Jo Fig.11: Adsorption kinetics model (a) relationship time and amount of chromium adsorbed per adsorbent (mg/g) (b) Pseudo-first-order kinetic model and (c) Pseudo-second-order kinetic model
of
ro
-p
re
lP
ur na
Jo
Table 1: Variables designed for the adsorption process.
Range and level variables
Adsorbent dosage(g) (A)
0.4
0.8
pH(B)
3
5
Contact time(min)(C)
30
90
₊α
of
α
1 8
Jo
ur na
lP
re
-p
−α
ro
Independent variables
150
Table 2: Full factorial design used for the chromium removal by adsorption process S.No
X1: Dosage(g) X2: pH
X3: Contact Y:Response time (min) Efficiency (%)
0.4
3 30
93.45
2
0.8
3 30
92.14
3 4 5 6 7 8 9 10 11 12 13 14
1 0.4 0.8 1 0.4 0.8 1 0.4 0.8 1 0.4 0.8
3 5 5 5 8 8 8 3 3 3 5 5
30 30 30 30 30 30 30 90 90 90 90 90
94.99 96.87 97.96 98.79 97.43 94.01 92.13 97.68 99.42 98.41 98.19 99.75
15 16 17 18 19 20 21 22 23 24 25 26 27
1 0.4 0.8 1 0.4 0.8 1 0.4 0.8 1 0.4 0.8 1
5 8 8 8 3 3 3 5 5 5 8 8 8
90 90 90 90 150 150 150 150 150 150 150 150 150
ro
-p
re lP
ur na
Jo
of
1
99.10 99.52 99.53 99.67 93.60 97.06 97.65 99.27 99.37 98.90 98.23 98.39 98.60
Table 3: Chemical composition of waste fish scale Chemical composition
Symbol
Percentage
1
Oxygen
O
37.9
2
Chlorine
Cl
35.4
3
Copper
Cu
8.7
4
Lead
Pb
6.2
5
Mercury
Hg
5.1
6
Zinc
Zn
7
Moisture
H2O
of
S.No
Jo
ur na
lP
re
-p
ro
5.9 1.8
Table 4: Analysis of variance (ANOVA) for response surface quadratic model Sum of
DF Mean square
pvalue
Prob>F
Model
114.7
9
12.74
6.85
0.0004
Significant
X1-Dosage
0.46
1
0.46
0.24
0.6275
Less significant
X2- Ph
12.21
1
12.21
6.49
0.0208
Significant
X3 -Cont. 27.87 time X12 0.036
2
27.87
14.81
0.0013
1
0.036
0.019
0.8922
of
Squares
F value
X22
25.14
1
25.14
13.36
0.0020
Significant
X32
35.44
1
35.44
18.84
0.004
Significant
X1X2
10.90
1
10.90
5.79
0.0278
Significant
X1X3
3.81
1
3.81
X2X3
1.81
1
1.81
Residual
31.99
17
Cor Total
146.69
26
Jo
re
-p
ro
Significant
2.02
0.1731
Insignificant
0.96
0.3410
In significant
lP
ur na
1.88
Significant
Table 5: Regression coefficient and the corresponding 95% CI High and Low
Standard
95% Cl
Low
High
Estimate
intercept
100.59
1
0.76
98.98
102.20
X1-Dosage
0.16
1
0.32
-0.52
0.84
1.04
X2-Ph
0.83
1
0.33
0.14
1.52
1.03
X3-Contact time X12
1.26
1
0.33
0.57
1.95
1.02
-0.088
1
0.64
-1.44
1.27
1.04
X22
-2.15
1
0.59
-3.39
-0.91
1.01
X32
-2.43
1
0.56
-3.61
-1.25
1.00
X1X2
-0.93
1
0.39
-1.74
-0.11
1.02
X1X3
0.55
1
0.39
-0.27
1.37
1.02
X2X3
0.39
1
0.39
-0.44
1.22
1.01
lP ur na Jo
VIF
-p
of
Factor
re
DF Error
95% CL
ro
Coefficient
Table 6: Constant values of Langmuir and Freundlich isotherm model
Freundlich parameters kf = 14.9107
KL= 91.001L/mg
1/n=0.008
3
RL = 0.00007325
-
4
R2=0.9998
R2= 0.9599
Jo
ur na
lP
re
-p
ro
2
of
Isotherms/Constant Langmuir parameters qmax= 18.3486mg/g 1
Table 7: Adsorption kinetic model parameters for chromium adsorption Pseudo 2nd order kinetic model
Initial Cr(III) qe(mg/g) ions (cal) concentration (mg/l) 18.8323
qe(mg/g)
K2
(exp)
(g/mg.min)
18.70116 0.0627
R2
qe(cal)
1
16.5576
1st
-p re lP ur na Jo
order
K1(min-1)
0.0953
ro
150mg/l
Pseudo model
kinetic R2
of
Kinetics
0.827
Table 8: Different bioadsorbent used for chromium (III) metal ions S.No
Bioadsorbent
Optimum parameters
1
Fish scale
IC= 250mg/50mL; 96% CT=120 min; D=20 g
Magsi et al., 2019
2
Chitosan
CT= 24hrs
Eladlani 2018
3
Jatropha curcas pH=5.5; D=8g/L; CT= 22.88mg/g L. 60mins
Gonçalves et al., 2017
4
Spirulina sp
Rezaei, et al., 2016
5
Residual brewing industry
6
Garden grass
IC=50mg/L; 90% (19.4 mg/g) Sulaymon et al., CT=180min; D=2g; pH= 2014 o 4; Temp=21 C
7
Ulva sp.
pH=4.5
8
Sugarcane bagasse
NA
9
Macrocystis pH=4 pyrifera and Undaria pinnatifida
10
Olive stone
70%
43.3%
re
lP
na
Vineyard pruning
12
Sargassum alga
et
al.,
Ferraz et al., 2015
-p
from IC=390 mg/L
Author
ro of
IC= 50mg/L; CT= 120 90.91 mg/g min; D= 0.1g; pH=5; Temp= 25oC
ur
Jo 11
Efficiency
2.89 mmol/g
Vijayaraghavan. and Joshi, 2014
73%
Ullah et al, 2013
0.77mmol/g and Cazón et al., 2012 0.74mmole/g
pH =4
5.18 mmole/g
Blázquez 2011
et
al.,
pH=4.2
12.45 mg/g
Karaoğlu 2010
et
al.,
sp. IC= 4.29mg/L; pH=3.5; 2.04 mg/g Temp=30oC
57
Silva et al., 2010
Waste fish scale IC=150 mg/L; CT= 90 99.75% min; D= 0.8g; pH =5; Temp= 21oC
Present research
Jo
ur
na
lP
re
-p
ro of
13
58