Вы находитесь на странице: 1из 13

Journal of Environmental Management 132 (2014) 237e249

Contents lists available at ScienceDirect

Journal of Environmental Management


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

Treatment and decolorization of biologically treated Palm Oil Mill Efuent (POME) using banana peel as novel biosorbent
Rae Rushdy Mohammed a, b, *, Mei Fong Chong a
a Department of Chemical and Environmental Engineering, Faculty of Engineering, University of Nottingham Malaysia campus, 43500 Semenyih, Selangor, Malaysia b Department of Chemical Industries, Mosul Technical Institute, Al-Majmoaa Al-Thaqaya, Mosul, Iraq

a r t i c l e i n f o
Article history: Received 27 June 2013 Received in revised form 11 November 2013 Accepted 18 November 2013 Available online 8 December 2013 Keywords: Palm Oil Mill Efuent Adsorption Biosorbent Banana peel

a b s t r a c t
Palm Oil Mill Efuent (POME) treatment has always been a topic of research in Malaysia. This efuent that is extremely rich in organic content needs to be properly treated to minimize environmental hazards before it is released into watercourses. The main aim of this work is to evaluate the potential of applying natural, chemically and thermally modied banana peel as sorbent for the treatment of biologically treated POME. Characteristics of these sorbents were analyzed with BET surface area and SEM. Batch adsorption studies were carried out to remove color, total suspended solids (TSS), chemical oxygen demand (COD), tannin and lignin, and biological oxygen demand (BOD) onto natural banana peel (NBP), methylated banana peel (MBP), and banana peel activated carbon (BPAC) respectively. The variables of pH, adsorbent dosage, and contact time were investigated in this study. Maximum percentage removal of color, TSS, COD, BOD, and tannin and lignin (95.96%, 100%, 100%, 97.41%, and 76.74% respectively) on BPAC were obtained at optimized pH of 2, contact time of 30 h and adsorbent dosage of 30 g/100 ml. The isotherm data were well described by the RedlichePeterson isotherm model with correlation coefcient of more than 0.99. Kinetic of adsorption was examined by Langergren pseudo rst order, pseudo second order, and second order. The pseudo second order was identied to be the governing mechanism with high correlation coefcient of more than 0.99. 2013 Elsevier Ltd. All rights reserved.

1. Introduction Malaysia currently accounts for 39% of world palm oil production and 44% of world exports. The oil palm planted area in 2011 reached 5.00 million hectares, an increase of 3.0% against 4.85 million hectares recorded in the previous year. Crude palm oil (CPO) production in 2011 increased by 11.3% to reach a record of 18.91 million tons (MPOC, 2013), which produced 30 million tons of Palm Oil Mill Efuent (POME) (Tengku et al., 2012). Palm oil processing is carried out in palm oil mills, where oil is extracted from a palm oil fruit bunch, and about 50% of the waste that results is POME. Despite high economical returns to the country, the industry also generates large amounts of waste in the form of empty fruit bunch (EFB) (23%), mesocarp ber (12%), shell (5%), and POME (60%) for every ton of fresh fruit bunch (FFB) processed in the mills

* Corresponding author. Department of Chemical Industries, Mosul Technical Institute, Al-Majmoaa Al-Thaqaya, Mosul, Iraq. Tel.: 964 7701622984. E-mail addresses: rae59.che@gmail.com, rafaa59@yahoo.com (R. R. Mohammed). 0301-4797/$ e see front matter 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.jenvman.2013.11.031

(Azhari et al., 2010). POME is a highly polluting wastewater that pollutes the environment if discharged directly into rivers due to its high chemical oxygen demand (COD), biochemical oxygen demand (BOD), phenol, and color concentrations. It is worth mentioning that no chemical is added during palm oil mill processing (Zahrim et al., 2009). Thus, the treatment of POME has gained interest from many researchers due to the abundant amount generated in the mills, and this treatment is an important issue for the minimization of water pollution. The high organic content, mainly oil and fatty acids, enables POME to support bacterial growth that reduces its polluting strength. Anaerobic process is the most suitable approach for its treatment (Azhari et al., 2010). Many technologies have been studied and applied for treating raw POME and biologically treated POME such as biological digestion (Chotwattanasak and Puetpaiboon, 2011), membrane technology (Ahmad et al., 2006), coagulation and occulation (Saifuddin and Dinara, 2011), adsorption (Igwe et al., 2010b), tertiary treatment (Shakila, 2008), and Fenton oxidation (Ooi, 2006) systems. A good review of treating POME is presented by Yeong et al. (2010). Every treatment has its own advantages and disadvantages. Membrane technology

238

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249 Table 1 Properties of biologically treated POME. Property pH Color TSS COD BOD5 Tannin and lignin Values 8.4 9900 1800 4700 1350 215 Standard discharge limits (Saifuddin and Dinara, 2011) 5e 9 e 400 100 e Units

has the highest removal efciency in COD but the treatment is very costly. As for carbon based adsorption, despite of its prolic use, granular activated carbon (GAC) is considered an expensive material. While many studies have been conducted about the use of activated carbon in treating various types of contaminants, there has been limited reporting of its application in treating POME. Also, there is little information in the eld about using biosorbents or activated carbon made from waste materials to treat POME. This study intends to ll the existing knowledge gaps. POME retains its color even after biological treatment, both aerobic and anaerobic. The color of the efuent is due to plant constituents such as lignin and phenolic compounds as well as repolymerization of coloring compounds after anaerobic treatment (Chanida and Poonsuk, 2011). The removal of color from efuents is one of the major environmental problems. In this concern, adsorption process has been found to be a more effective method for the treatment of dye containing wastewater. The most efcient and commonly used adsorbent is commercially available activated carbon which is expensive and has regeneration problems. Recent investigations focused on effectiveness of low cost adsorbents like pearl millet husk, neam leaf powder, coconut husk, wheat straw, wood, peat, banana pith, and agricultural waste in the removal of dyes from wastewater efuent (Verma and Mishra, 2010). However, there is still scarcity of information in the literature on the use of low cost adsorbents for decolorization of POME. As one of the most consumed fruits in the world, banana is a very common fruit. The main banana residue is the fruit peel, which accounts for 30e40% of the total fruit weight. Preliminary investigations show that several tons of banana peels are produced daily in marketplaces and household garbage, creating an environmental nuisance and disposal problem. Various chemical groups exist on the banana peel surface, including carboxyl, hydroxyl and amide groups, which have been extensively proven to play a critical role in the biosorption processes (e.g. enhancing biosorption capacity and shortening stable time) (Cong et al., 2012). In this study, efciencies of removal of color, TSS, tannin and lignin, BOD, and COD from the nal biologically treated POME were investigated using 3 types of adsorbents: natural banana peel (NBP), methylated banana peel (MBP), and banana peel activated carbon (BPAC). The effects of various parameters such as agitation time, pH, and adsorbent dosage were investigated in batch experiments. Equilibrium isotherms were analyzed by using the Langmuir, Freundlich, Redlich, and Sips models. The adsorption kinetic was determined by tting with pseudo-rst-order, pseudo second order, and second order adsorption kinetic models. Determination of the isotherm and kinetic concepts provided a sound basis for the process of designing an adsorption unit for POME nal polishing to achieve optimal treatment results. Thus, banana peel can be used in removing color from biologically treated POME as the nal polishing step before discharge. 2. Materials and methods 2.1. Biologically treated POME The efuent was collected in plastic containers from the nal pond efuent of a palm oil mill in Dengkil, Selangor, Malaysia. The containers were properly washed and rinsed with the efuent before collection to avoid contamination and dilution. Containers were then brought back to laboratory and stored in refrigerator at temperature of 4  C for tests and analysis. COD, pH, TSS, tannin and lignin, BOD, and color units were determined. The characteristics of the biological POME sample obtained are summarized in Table 1. It has been observed that the biologically treated POME exceeded the standard discharge limit of Environmental Quality Act

PtCo/l mg/l mg/l mg/l mg/l

(EQA) 1974, Department of Environment (DOE), Malaysia (Saifuddin and Dinara, 2011). The nal treated efuent still contained high concentrations of COD, color, BOD and TSS. On the other hand, the pH indicated that the efuent was alkaline. All chemicals used in this study were of analytical grade supplied by Aldrich Chemicals. All solutions used in this study were diluted with distilled water as required. 2.2. Collection and preparation of adsorbents 2.2.1. Natural banana peel (NBP) Mature banana with yellow peel was collected as solid waste. The collected material was then washed three times with tap water and three times with distilled water to remove external dirt. The washed material was cut into small pieces (1e2 cm) then dried in a hot air oven (Memmert Universal oven Model UFE 600 - Germany) at 80  C until it reached a constant weight, which was accomplished after 48 h. In the nal stage, the material was grounded by using Retsch Cutting Mill SM 100 (Germany) with mesh size 0.2, and screened by using ELE international laboratory sieve shaker (USA) with mesh size of 300e425 mm. 2.2.2. Methylated banana peel (MBP) Modication of the carbonyl groups on the surface of the banana peel (esterication) was achieved by using acidic methanol. 9 g from the previously prepared NBP was suspended in 633 ml of 99.9% methanol to which 5.4 ml of concentrated hydrochloric acid 37% was added to give a nal concentration of 0.1 M. Then the solution was heated at 60  C and stirred continuously for 48 h by using digital orbital shaker (Heidolph unimax 1010, Germany). The solid material was then separated and washed three times with deionized water at (20  C) in order to halt the esterication reaction. The material was then dried in the oven at 100  C for a period of 8 h (Jamil et al., 2008). 2.2.3. Banana peel activated carbon (BPAC) The NBP was used for BPAC preparation. It was prepared by a carbonized temperature of 500  C for 1 h under a nitrogen atmosphere in a tubular furnace (Carbolite CTF 12/100/900, United Kingdom) and then was left to cool to room temperature. The carbonized material was then subjected to potassium hydroxide (KOH) activation. The material was agitated by using (cermaic stirring; IKA 3581000; Vernon Hills, IL, USA) in KOH (45% aqueous solution) at a ratio of 1:4 carbonized material to KOH weight by weight basis. After 12 h of agitation, the carbonized slurry was left for 24 h at room temperature. The sample was then dried at 110  C for another 24 h. The dried sample was then activated in nitrogen atmosphere using (Carbolite CTF 12/100/900, United Kingdom) with temperature maintained at 700  C for 1 h before cooling. After cooling for 12 h, the activated carbon was washed 3 times with 0.2 N hydrochloric acid (HCl). The washing was completed with hot water until the pH became neutral, and nally with cold water to

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

239

remove the excess KOH compounds. The washed samples were dried at 110  C for 8 h to get the nal product (Mopoung, 2008). 2.3. Adsorption isotherm experiments Adsorption isotherm experiments were carried out in 250 ml conical asks into which 15 g of adsorbent and 100 ml of biologically treated POME were added with different concentrations respectively. The desired concentrations were achieved by diluting the biologically treated POME with distilled water. The samples were then shaken at 200 rpm for 40 h by using digital orbital shaker (Heidolph unimax 1010, Germany). At the end of the adsorption period, the solution was centrifuged for 5 min at 3000 rpm and then the concentrations of the residual color, TSS, and COD were determined. The adsorption capacity of these parameters on the adsorbent was calculated from the mass balance equation as follows

qe Co Ce V =M

(1)

where qe is the amount of constituent adsorbed per unit mass of adsorbent at equilibrium (mg/g), Co and Ce are the initial and equilibrium liquid-phase concentrations of solution (mg/L) respectively, V is the volume of efuent solution (L), and M is the mass of adsorbent sample used (g). 2.4. Adsorption experiment In each adsorption experiment, 100 ml of efuent solution was added to different amounts of adsorbent (5, 10, 15, 20, 25 and 30 g) respectively in closed asks and the samples were stirred at 200 rpm in a rotary orbital shaker (Cermaic Stirring; IKA 3581000; Vernon Hills, IL, USA) at room temperature for a period of time ranging from 2 to 36 h. Shaking step was done after noting down the initial pH of the solutions (2, 5, 7, 8.4, and 12). The initial pH of each solution was adjusted to the required value by adding either 1 M HCl or 1 M sodium hydroxide (NaOH) solution. Samples were withdrawn from each of the asks at predetermined time intervals. The adsorbents were separated from the solution by centrifugation (Eppendorf 5430, Hamberg, Germany) at 3000 rpm for 5 min. Color, TSS, COD, BOD, and tannin and lignin were measured to determine the concentration of the residue. The percentage of removal (%R) of each parameter and the amounts adsorbed by the adsorbents were calculated by the following equations:

APHA Standard method 8006 using HACH spectrophotometer (DR 2800, Loveland, CO) by placing 10 ml of sample in a specially designed square quartz sample cell. The color analysis was carried out by following platinum-cobalt standard method 8025 and the color was detected at a wavelength of 455 nm using HACH spectrophotometer (DR 2800, Loveland, CO) by placing 10 ml of sample in the same previous sample cell. Tyrosine method 8193 was used with the aid of HACH spectrophotometer (DR 2800, Loveland, CO) to measure the concentration of tannin and lignin in the sample. After adding 0.5 ml of tannin-lignin reagent and 5.0 ml of sodium carbonate solution to 25 ml of the sample, 10 ml of the mixture was poured into square sample cell and the measured result was in mg/l tannins. The BOD tests were carried out using standard procedure APHA Standard method with a dissolved oxygen (DO) probe (YSI model 5100, USA). Seeds and nutrient buffer pillows were added to the samples (1.5e3.5 ml). Samples were diluted 15 times followed by aeration to make sure enough dissolved oxygen will remain after ve days of incubation. 300 ml standard BOD bottles were used in these tests which were incubated at 20  C in a dark incubator (Memmert Model IN110 e Germany). After ve days, the differences in oxygen consumption were measured by using a DO meter (YSI model 5100, USA) to estimate BOD5. Solutions pH were measured with a HACH sension1 pH meter using a combined glass electrode. Finally, surface morphology and BET surface areas of NBP, MBP and BPAC were studied using Quanta 400 Field Emission Scanning Electron Microscope (FESEM), and N2 adsorption isotherm by Nano Porosity System (Micrometrics ASAP 2020, USA) respectively. 3. Results and discussion In this study, the adsorption behavior of biologically treated POME onto 3 different types of biosorbents, namely NBP, MBP and BPAC prepared from banana peel was investigated. The BET surface area of NBP was found to be 24.2572 m2/g. This BET surface area was higher than that of chitosan akes and pine bark (Ngah and Fatinathan, 2006; Vzquez et al., 2007) but lower than that of palm oil fruit shell (Hossain et al., 2012). On the other hand, the surface area was low compared to chemically (MBP), (168.3648 m2/ g), and thermally (BPAC), (875.2914 m2/g), treated banana peel. Nevertheless, it could be considered a good alternative as it was produced from non-treated, low cost material using a simple process. It is evident that specic surface area of banana peel biosorbent is fully dependent on the preparation method. In general, acidic methylation and alkali treatment of the NBP to produce MBP and BPAC respectively could alter its lattice structure effectively compared to the raw material. These treatments led to remarkable increase in the specic area and the major change achieved in the materials structure promoted considerably its capacity for all contaminants removal compared to original untreated peel. 3.1. Performance study In the present study the adsorption performance is measured on the basis of the percentage of removal for the parameters of color, TSS, and COD from biologically treated POME as in Equation (2). The inuence of operating parameters in terms of biosorbent dosage, contact time, and initial pH on the adsorption performance was investigated for all types of biosorbents. Fig. 1 shows the effect of adsorbent dosage at room temperature (25  C) by varying the sorbent amounts from 5 to 30 g with varying contact times. Fig. 1 shows the effects of NBP, MBP and BPAC dosage on the percentage of removal of color, TSS, and COD. It is obvious that, regardless of the type of biosorbent used, the percentage of removal increases

%R fCo Ct =Co g 100 qt Co Ct V =M

(2) (3)

Here, Ct is the concentration of the solution at any time (mg/L) and qt is the amount adsorbed at any time for each parameter investigated (mg/g). Three repeatable experiments were tested and the average values with their standard deviations were recorded and used for discussion. 2.5. Analytical methods COD measurement was carried by APHA Standard method 8000 using the colorimetric method. 2 ml of sample was added into COD vial (HR 20e1500 mg/L) and digested for 2 h by using HACH digester (DRB200, Loveland, CO). The COD concentration was then measured with a HACH spectrophotometer (DR2800, Loveland, CO). Total suspended solids were also measured by following the

240 R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

Fig. 1. Effect of dosage on the percentage removal of (i) color, (ii) TSS, and (iii) COD for (a) NBP, (b) MBP and (c) BPAC respectively with varying contact time at pH 8.4.

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

241

rapidly with increase in the amount of adsorbent due to greater availability of surface area for adsorption. This leads to the introduction of more binding sites for adsorption. A signicant increase in uptake was observed when the dosage was increased from 5 to 25 g/100 ml. Any further addition of adsorbent beyond this amount did not cause signicant change in adsorption performance. This may be due to the overlapping of adsorption sites as a result of overcrowding of adsorbent particles (Goud et al., 2005; Verma and Mishra, 2008). For all types of biosorbents, maximum removal of all studied contaminations was obtained at the adsorbent dosage of 30 g/100 ml. It can be seen from Figs. 1 and 2 that by increasing the contact time from 2 h to 36 h the removal percentage increases until adsorption equilibrium is reached. It also illustrates that the contaminant removal was rapid in the rst 24 h during which nearly 70e80% of the total uptake appeared to have been adsorbed depending upon the adsorption ability of different biosorbents. This can be attributed to the availability of sites for the sorbate. In addition, a very high adsorption driving force at the beginning resulted in a higher adsorption rate. After the initial period, slower adsorption may be attributed to the slower diffusion of molecules into the interior pores of the adsorbent (Shavandi et al., 2012), and the molecules subsequently occupy the positions within the adsorbent framework. This observation is in support of the ndings reported by several authors (Goud et al., 2005; Shavandi et al., 2012; Zahangir et al., 2009).

Fig. 3. Effect of pH on the percentage removal of color, TSS and COD on 30 g/100 ml MBP adsorbent at 24 h.

Suspended and dissolved particles in water inuence color. Dissolved organic matter, such as humus, peat, or decaying plant matter can produce a yellow or brown color. Concentrations of naturally dissolved organic acids such as tannins and lignins may also have an effect by giving water a tea color. Tannins that are yellow to black are the most abundant kind found in POME and can have a great inuence on its color, as well as a musty smell. The brown coloring comes from tannins leaching into runoff water from the manufacturing process of palm oil (Clean Water team, 2013). From the previous facts, it is obvious that color removal is strongly dependent on the removal of total suspended solids and tannin and lignin. This is revealed by Fig. 1(i), (ii), and 2(a). LAB Fit curve tting software V7.2.48 was used to establish a new correlation between the adsorption capacities of color with TSS and tannin and lignin on BPAC. This correlation is represented by:

Y 37:18X1 *X2 0:0468

(4)

where Y, X1 and X2 are the adsorption capacity of color, TSS and tannin and lignin respectively. The statistical data for the predicted correlation are (Correlation Coefcient: R2 0.9938680E00, ChiSq. 0.145292E00, Red. ChiSq. 0.363231E-01 > P(Red. ChiSq.) 0.997 and Average absolute residual 0.138289E00). This correlation may offer a basis for the estimation of color removal as a function of total suspended solids and tannin and lignin removal from POME. This correlation is also indicates that TSS and tannin and lignin contents of POME are responsible for its color. The effect of variation of pH on adsorption performance in terms of color, TSS, and COD onto MBP at constant time of 24 h and adsorbent dosage of 30 g/100 ml are shown in Fig. 3. As other biosorbents of NBP and BPAC follow a similar trend as that in Fig. 3

Table 2 Comparison of biosorption capacity of color, TSS and COD on NBP, MBP and BPAC adsorbent dose (5e30 g/100 ml), contact time (2e36 h) and pH (2e12). Parameters (% removal) Biosorbent NBP Min. Fig. 2. Effect of contact time on the percentage removal of (a) tannin and lignin with varying dosage and (b) BOD with the dosage of 5 g/100 ml on BPAC with constant pH of 8.4. Color TSS COD 18.182 5.556 70.213 Max. 61.111 90.278 99.468 MBP Min. 41.414 5.556 80.850 Max. 73.737 91.667 w100 BPAC Min. 64.646 50.000 84.04 Max. 96.464 w100 w100

242

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

at different levels of intensity, the removal efciency at minimum and maximum percentages for various contaminants is tabulated in Table 2. The maximum percentages of removal in Table 2 are for 30 g/100 ml adsorbent dose, 30 h contact time and pH 2. The minimum values are for 5 g/100 ml, 2 h contact time and pH 12. A deterioration of adsorption performance as pH increased was observed for all types of biosorbent with the maximum percentage of removal recorded at pH 2. The same trend for NBP was found in another study where Khan et al. (2012) found that banana peel

adsorbent was effective in the adsorption of Reactive Yellow 15 from aqueous solution and maximum adsorption occurred at pH 2. The most possible explanation may be related to the presence of excess H ions accelerating the removal of the contaminants with the anion OH- in the solution. It is also possible that the surface properties of the adsorbent depend on the pH of the solution. Higher uptakes obtained at lower pH may be due to the electrostatic attractions between negatively charged functional groups located on the contaminants and positively charged adsorbent surface. Hydrogen ion also acts as a bridging ligand between the

Fig. 4. Adsorption isotherm of (a) color, (b) TSS, and (c) COD on NBP.

Fig. 5. Adsorption isotherm of (a) color, (b) TSS, and (c) COD on MBP.

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

243

adsorbent wall and the contaminants molecules (Zawani et al., 2009). Malik (2004) and Mohamed (2003) reported that at low pH region the surface of the sorbent will be largely protonated. The positive ions (H) provide an electrostatic attraction between the bers surface and the contaminants molecules leading to maximum adsorption. On the other hand, at pH above 2 (i.e., pH range of 3e12) the degree of protonation of the surface of the bers will be less, which results in a decrease in diffusion and adsorption due to electrostatic repulsion. Furthermore, lower adsorption of the contaminants in alkaline medium can also be attributed to the

Fig. 6. Adsorption isotherm of (a) color, (b) TSS, and (c) COD on BPAC.

competition between excess hydroxide ions (OH) and the anionic contaminant molecules for the adsorption sites (Khan et al., 2012). In all, it is clear that NBP, MBP, and BPAC were capable of reducing the concentrations of color, TSS, COD, BOD, and tannin and lignin from biologically treated (POME). The adsorptive removal of these contaminants could be explained based on the modes of sorption of materials or contaminants on biosorbents. The adsorptive removal of contaminants may be attributed to two main terms; intrinsic adsorption and coulombic interaction. The coulombic term results from the electrostatic energy of interactions between the adsorbents and adsorbates. It can also be observed in the adsorption of cationic species versus anionic species on adsorbents. The intrinsic adsorption of the materials is determined by their surface areas. Moreover, both factors can interact, thereby inuencing the adsorption capacity. It has been reported that the surface area has a great effect on the sorption capacities of adsorbents (Igwe et al., 2010a). Thus, increase in surface area increases sorption capacity. The trend of this sorption capacity can be put as COD > BOD > TSS > tannin and lignin > color (Figs. 1 and 2 and Table 2). This trend can be explained based on solubility and diffusion processes. COD and BOD are dened as the amount of dissolved oxygen needed to break down the carbonaceous component of the waste (Igwe et al., 2010a). This means that both COD and BOD involve dissolved components of the efuent. Also, color is a consequence of dissolved components and TSS of the efuent. Diffusion takes place before adsorption of the contaminants because dissolved contaminants diffuse faster than suspended particles, hence they will be more adsorbed. Therefore, in our opinion, this is the reason for the sorption trend observed. The pH could be considered as resulting from the interactions of the other parameters. That is, changes in pH of the efuent are dependent on the changes in the hydrogen ion concentration which in turn is dependent on the dissolved components of the efuent. It was found that BPAC biosorbent exhibited the highest performance in terms of color, TSS, COD, BOD, and tannin and lignin removal efciency followed by MBP and NBP. This may be attributed to the carbonization followed by chemical (KOH) activation of the banana peel. Chemical activation of carbons is a very common method for obtaining activated carbons with very high surface areas. KOH is one of the most effective agents employed for organic materials. KOH may be more selective in the activation process, causing a more localized reaction with the carbon precursor and is more effective for highly ordered materials (Mopoung, 2008). The effectiveness of KOH activation relative to either physical activation methods or activation by other chemical agents can be attributed to the ability of potassium (K) to easily form intercalation compounds with carbon. In addition, the potassium oxide (K2O) formed during the process of KOH activation, in-situ, can easily inltrate into the pores. K2O is reduced to K by carbon resulting in carbon gasication with a subsequent emission of carbon dioxide (CO2) leading to the formation of pores. Also K atoms that intercalate into the lamella of the carbon crystallites widen the space between the adjacent carbon layers (intercalation phenomenon), resulting in an increase in the value of specic surface area (Yong et al., 2007). The carboxyl groups on banana surface are responsible to some extent for the binding of sorbate ions. This means that increasing the number of carboxylate ligands in the biomass can enhance the sorbate binding capacity. Some groups from cellulose, pectin, and hemicelluloses, which are major constituents of banana peel, can be modied to carboxylate ligands by treating the biomass with KOH, thereby increasing the sorbate-binding ability of the biomass (Baig et al., 1999). In general, acidic methylation, alkali, and thermal treatment of the NBP to produce MBP and BPAC respectively could alter their lattice structure more effectively compared to the raw material.

244

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

Table 3 Adsorption isotherm parameters, correlation coefcient and relative average error for Langmuir, Freundlich, RedlichePeterson and Sips isotherm for adsorption of Color, TSS and COD using NBP, MBP and BPAC. Biosorbent Parameter NBP Color TSS 28.409 1.983 103 91.19 102 0.087318 21.9 102 46.2 102 1.85 0.937 0.088169 39.27 102 53.42 102 52.07 102 0.9750 0.086435 53.07 102 52.01 102 1.248 104 0.9449 0.090476 COD 67.563 9.113 104 96.71 102 0.033989 18,9 102 39.1 102 1.578 0.9939 0.025039 12.16 102 6.312 102 55.28 102 0.9933 0.022765 24.88 102 73.11 102 1.502 103 0.9914 0.023493 MBP Color 87.719 3 104 98.04 102 0.132826 25 102 46.7 102 1.745 0.9776 0.048738 2.555 102 1.899 105 1.279 0.9950 0.04144 1.754 102 1.081 1.802 104 0.9939 0.045303 TSS 47.17 1.206 103 89.72 102 0.122038 31.5 102 50.5 102 1.7 0.9908 0.035662 133.5 286.6 39.98 102 0.9979 0.034923 37.14 102 64.67 102 1.107 103 0.9939 0.03716 COD 147.059 3.717 104 87.43 102 0.039245 36.4 102 16.3 102 1.26 0.9975 0.022178 987.6 6970 18.63 102 0.9980 0.020235 18.26 102 76.67 102 3.253 104 0.9981 0.019674 BPAC Color 135.135 1.675 103 99.7 102 0.03911 5.69 102 662.7 102 2.837 0.9532 0.062763 26.70 102 3.352 103 93.94 102 0.9986 0.036194 57.15 102 84.31 102 3.948 103 0.9952 0.038162 TSS 62.893 3.721 103 98.3 102 0.055281 13 102 110.6 102 1.719 0.9602 0.074782 19.94 102 8.486 104 1.197 0.9994 0.043963 7.565 102 1.251 1.438 103 0.9969 0.047905 COD 200 1.91 103 92.6 102 0.102418 1 101 187.1 102 1.613 0.9935 0.029085 85.74 102 11.38 102 57.07 102 0.9982 0.023811 1.354 71.65 102 3.988 103 0.9951 0.025001

Langmuir isotherm qm (mg/g) 49.751 b (L/mg) 3.559 104 2 R 98.84 102 Relative av. error 0.027004 RL 22.1 102 Freundlich isotherm Kf 28.5 102 n 1.805 0.9624 R2 Relative av. error 0.062819 RedlichePeterson Isotherm KRP 1.408 102 aRP 1.500 105 b 1.321 R2 0.9945 Relative av. error 0.017747 Sips Isotherm KS 3.489 103 bS 1.234 aS 8.083 105 R2 0.9927 Relative av. error 0.020385

These treatments led to remarkable increase in the specic area, and the major change achieved in the materials structure considerably promoted its capacity for all contaminants removal compared to the original untreated peel. 3.2. Equilibrium isotherms In this study equilibrium isotherm equations are used to describe the experimental sorption data. The equation parameters and the underlying thermodynamic assumptions of these equilibrium models often provide some insight into the sorption mechanism and the surface properties as well as afnity of the sorbent. The most common isotherms applied in solid/liquid systems are the theoretical equilibrium isotherm, the Langmuir, which is the best known and most often used isotherm for the sorption of a solute from a liquid solution; the Freundlich, which is the earliest known relationship describing the adsorption equation; the Redliche Peterson and the Sips, which are the isotherms containing three parameters. These equilibrium isotherm models were used to analyze the adsorption behavior in this study. Table S1 supplementary data shows a summary of the models used. The equilibrium isotherm data were tted with Langmuir and Freundlich models by using linear regression method. Non-linear regression was used for data tting on RedlichePeterson and Sips models with the aid of Graphpad Prism version 6 package. The degree of tness for all models was determined by least squares error analysis. From Table S1 supplementary data, qm is the maximum uptake capacity under the given conditions (mg/g), b is the equilibrium constant related to free energy of adsorption (L/mg), Kf (mg/g)(L/ mg)1/n and n are the Freundlich constants characteristic of the system and the indicators of the adsorption capacity and adsorption intensity respectively, KRP (L/g) and aRP (L/mg)(1/b) are Redliche Peterson isotherm constants, Ks (mg/g)(L/mg)bs and as (L/mg)bs are Sips isotherm constants, b and bs are the isotherm exponents. The comparison of the experimental and estimated data by Langmuir, Freundlich, RedlichePeterson, and Sips models for the parameters of color, TSS, and COD with different biosorbents of NBP, MBP and BPAC at 25  C are presented in Figs. 4e6 with the

corresponding goodness of t in Table 3. Generally, both twoparameter and three-parameter models show good tness with experimental data for adsorption of most parameters. This can be attributed to the presence of both heterogenous and homogenous surfaces on the banana peel (Febrianto et al., 2009). However, based on values of mean residual square (R2) and relative average error as criteria for goodness of t, RedlichePeterson isotherm provides a better correlation between the theoretical and experimental data for the whole concentration range for all types of biosorbent compared to the other models. This is expected as the form of the RedlichePeterson equation includes features of the Langmuir and Freundlich isotherms. So it may be used to represent adsorption equilibrium over a wide concentration range of adsorbate. The values of adsorption capacity qm for adsorption of color, TSS and COD onto BPAC are higher than for both MBP and NBP. The higher adsorption capacity for all parameters for the BPAC in comparison to the other studied adsorbents may be attributed to the higher pore fraction capable of adsorbing the contamination molecules and is in-line with the trend observed from the previous section of performance study. However, heterogeneity of the carbon surface and the wide range of pore sizes and surface properties extremely complicate analysis of the observed behavior. 3.3. Adsorption kinetics The experimental sorbate uptake rates onto the biosorbents were investigated by using kinetic models in this study. The kinetics of adsorption based on the overall adsorption rate by the adsorbents were analyzed by the Pseudo-rst-order, pseudo-second-order, and second order kinetic models as shown in Table S2 supplementary data. In Table S2 supplementary data, qt is the amount adsorbed at time t (mg/g); K1 is the rate constant of rst order adsorption (L/ min), K2 and K are the rate constants of pseudo second order and second order-adsorption respectively (g/mg.hr). Figs. 7e9 show that for the pseudo rst order kinetic model there is a deviation from the experimental data and the data were tted with a poor correlation coefcient and high relative average

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

245

error for most parameters (Table 4). This indicates that the rate of removal of color, TSS, and COD onto banana peel does not follow the pseudo-rst-order sorption rate expression of Lagergren. The pseudo-second-order equation is based on the sorption capacity of the solid phase. It predicts the behavior over the whole range of data. Furthermore, it is in agreement with chemisorption being the rate controlling step. Model parameters K2 and qe values

were determined from the slope and intercept of the plots of t/qt against t. The values of the parameters, correlation coefcients and relative average errors are also presented in Table 4. The correlation coefcients of all examined data were found to be very high (R2 > 0.99) and the relative average errors were very low. Also the calculated values of equilibrium sorption capacity, qe, are very much in agreement with experimental data for all studied systems. This shows that the model can be applied for the entire adsorption process and conrms that the sorption of color, TSS, COD, BOD, and

Fig. 7. Comparison of different kinetic models for (a) color, (b)TSS and (c) COD on NBP biosorbent.

Fig. 8. Comparison of different kinetic models for (a) color, (b)TSS and (c) COD on MBP biosorbent.

246

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

tannin and lignin onto NBP, MBP and BPAC follow the pseudosecond-order kinetic mechanism. Figs. 7e9 also show the kinetic modeling of color, TSS, and COD adsorption by the second order model equation. Correlation coefcients, relative average errors and the second order model rate K for all studied systems are listed in Table 4. The correlation coefcients, R2, of the second-order kinetic were found to be relatively low compared to the pseudo second order model, but still better than those of the rst order kinetic model. Relative average

errors were higher than that of the pseudo second-order model, but still lower than those of the pseudo rst-order model. The kinetics of sorption processes are concerned with forces between sites and adsorbate molecules, and this forms an important area of surface chemistry. Banana peel surface is cellulose based and contains carboxylic and amine groups. Also the esterication process of the banana peel introduces specic functional groups on the bril surface by which signicant improvement of hydrophobicity may be obtained. On contact with water and depending on pH, these groups become negatively charged and are likely sites for chemical reaction to take place on the banana peel surface. According to Ho and McKay (1999), pseudo-second order model is based on the assumption that the rate-limiting step may be chemical sorption or chemisorption involving valency forces through sharing or exchange of electrons between sorbent and sorbate. The adsorption system obeys the pseudo second-order kinetic model for the entire adsorption period and thus supports the assumption behind the model that the adsorption is due to chemisorption. 3.4. Scanning Electron Microscopy (SEM) The surface physical morphology of NBP, MBP, and BPAC as observed in Fig. 10 show progressive changes in the surface of the particles. Fig. 10(a) reveals that microporous structures, heterogenous, rough surface with crater-like pores exist in banana peel. The NBP particles are of irregular shape and their surface exhibits a micro-rough texture, which can promote the adherence of adsorbates. Fig. 10(b) shows that the treatment with methanol and acid has changed the surface morphology of NBP. The surface of MBP has more irregular and porous structure than that of NBP, and therefore explains the higher BET surface area with higher adsorption capacity. Fig. 10(b) also shows that the pores within the particles are more homogenous. Also, this slightly rough surface provides suitable binding sites for adsorbate molecules. It is obvious from Fig. 10(C1) that the BPAC activated by KOH lost its original cellular structure and looked broken. The KOH reagent is a strong base. It is able to interact with carbon atoms and thus catalyze the dehydrogenation and oxidation reactions, leading to the increment of tar evolution and development of porosity (Hsu and Teng, 2000). White spheres and some uffy materials can be seen in the pores of BPAC (Fig. 10(C2)). The white spheres and uffy materials may be due to the presence of K2O, potassium carbonate (K2CO3), or K residues. It may be argued that during the KOH activation process, various reactions can take place with such products as hydrogen gas (H2), K, K2CO3, and K2O (Babel and Jurewicz, 2004). At high activation temperature, the formation of K2O is thermodynamically the most stable. The high KOH ratio of 4 to 1 in samples containing a large amount of K may cause the production of more K2CO3 and K2O during pyrolysis. It appears that larger amounts of K2CO3 cause larger pore size and structural deformation (Fig. 10(C3)). This proposed mechanism causes BPAC to have a higher performance than NBP and MBP (Yupeng et al., 2002). These ndings are in agreement with the ndings of Mopoung (2008). 3.5. Optimum conditions for treatment and decolorization of biologically treated POME Treatment of POME requires a sound and efcient system to face the current challenges. With the present situation where there are some mills still failing to comply with the DOE standard discharge limit even after they have applied the available treatment system, it is believed that adsorption technology will be able to further polish

Fig. 9. Comparison of different kinetic models for (a) color, (b)TSS and (c) COD on BPAC biosorbent.

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

247

Table 4 Parameters, correlation coefcient and relative average error of kinetic models obtained by using the linear methods for adsorption of color, TSS, COD, Tannin & lignin and BOD on NBP, MBP and BPAC. Adsorbent Adsorbate qe(exp.) (mg/g) Pseudo rst-order model K1 (1/hr) R2 NBP Color 65.0 TSS 19.0 COD 85.0 Color 94.0 TSS 13.0 COD 85.0 Colour 176.0 TSS 28.5 COD 91.5 Tannin & lignin 1.4 BOD 27.0 0.2633 0.1275 0.2829 0.4505 0.2195 0.4501 0.2183 0.1854 0.2472 0.1153 0.1257 0.9550 0.9960 0.6710 0.9150 0.9710 0.8090 0.7742 0.7166 0.7920 0.9788 0.9674 Pseudo second-order model R2 0.9983 0.9940 0.9996 w1 0.9940 w1 0.9999 0.9988 0.9995 0.9988 0.9991 Second-order model Relative av. Error 0.145756 0.339175 0.021715 0.023886 0.304766 0.038508 0.087279 0.097595 0.027888 0.07071 0. 09533

Relative av. Error K2 (g/mg/hr) 0.056534 0.06474 0.102363 0.073979 0.114048 0.074324 0.148375 0.153404 0.15139 0.17692 0.10591 6.762 10 2.913 103 11.710 103 39.400 103 19.700 103 47.200 103 6.436 103 23.600 103 26.400 103 90.400 103 4.965 103
3

Relative av. Error K (g/mg/hr) R2 0.033199 0.027864 0.016083 0.006587 0.050348 0.00709 0.012703 0.035884 0.018225 0.010357 0.070284 0.0314 0.0190 0.0156 0.0911 0.0910 0.8810 0.0663 0.0871 0.0861 0.4209 0.0609 0.9265 0.9709 0.9931 0.9327 0.9568 0.9523 0.9034 0.9135 0.8577 0.9664 0.7092

MBP

BPAC

the biologically treated POME in a more benecial way. For all studied biosorbents, the optimum condition was found at contact time of 24 h, dosage of 30 g/100 ml, and pH of 2. The BPAC was found to be the optimum biosorbent as compared to NBP and MBP. Under optimum conditions, the color of biologically treated POME (dark brown) was signicantly different after adsorption process. The color after adsorption by NBP and MBP turned to light and very light yellow respectively. The color degree of biologically treated POME after undergoing BPAC adsorption treatment was further reduced, as more than 96.46% of its initial value was removed, and it was found to be colorless to the naked eye. The nal color concentration at the optimum conditions of this work was 350 PtCo/L after treatment with BPAC adsorption compared to a nal color concentration of 1350 PtCo/L found by Kutty et al. (2011) after treating POME with microwave incinerated rice husk ash adsorption. The test results for TSS removal showed that the maximum extent of removal was approximately 90% after treating with NBP. This is equivalent to a residual concentration of 180 mg/L which is much lower than the standard discharge limit. This removal percentage is slightly higher when treating with MBP where it reaches 91.667% of the initial value, while the nal TSS concentration after BPAC adsorption treatment was close to 0 mg/L. COD shows a reduction by 99% after adsorption with NBP. The MBP and BPAC treatments brought down the COD value nearly to 0 mg/L from the original COD value of biologically treated POME. For BOD reduction, a similar trend to the COD reduction was obtained from each type of biosorbent. However, it can be seen that NBP treatment process reduced the BOD percentage by more than 90%, which is well below the allowable limit set by the Malaysian Department of Environment, which is 100 mg/L. Similarly, a BOD reduction to near 0 mg/L was observed for both MBP and BPAC treatment processes. Adsorption processes were expected to reduce tannin and lignin content of biologically treated POME effectively. Results revealed that BPAC adsorption processes were capable of reducing tannin and lignin by up to 76.744% (as compared to initial level). Due to the fact that there is no chemical addition during palm oil processing, the color of this efuent is due to plant constituents such as lignin and phenolics as well as repolymerizsation of coloring compounds after anaerobic treatment (Zahrim et al., 2009). Considering the results of this research, it is believed that decolorization of biologically treated POME is a function of its TSS and tannin and lignin concentrations. In addition, with the nal pH adjustment to 7, the resulting biologically treated POME using this technology is in compliance with standard discharge regulations and the resulting high-quality treated water can be recycled back to the plant to be reused.

4. Conclusion Research was carried out to explore banana peels as a high capacity, economically viable and low cost adsorbent for treating biologically treated POME. The experimental results showed that NBP has the ability to reduce the concentrations of biologically treated POME to meet the POME discharge standard. Results showed that modication of NBP by acidied methanol or KOH and thermal treatment can further improve the sorption removal for color, TSS, and COD. The optimum adsorption capacity of NBP was found at 97 mg/g for color, 25 mg/g for TSS, and 90.5 mg/g for COD removals respectively. The MBP has optimum adsorption capacity of 137.5 mg/g, 28.5 mg/g and 93 mg/g for color, TSS, and COD removals respectively, while BPAC has 184.5 mg/g, 34.5 mg/g, 94 mg/ g, 26.76 mg/g and 1.4 mg/g for color, TSS, COD, BOD, and tannin and lignin removals respectively. In the present study, the optimum adsorption conditions were found at dosage of 30 g of adsorbent/ 100 ml of sample and after 24 h contact time. The pH played an obvious effect on the sorbate adsorption capacity onto all types of banana peel biosorbents. The decrease in the solution pH led to a signicant increase in the adsorption capacities of all parameters on the banana peel biosorbents with maximum adsorption capacity occurring at acidic pH of 2, although, the adsorption efciency is still high at the original efuent pH (8.4). Four adsorption isotherm models were studied. The sorption data of color, TSS, and COD on banana peel biosorbents tted into two-parameter isotherm models (Langmuir and Freundlich) and three-parameter isotherm models (RedlichePeterson and Sips). An excellent prediction in all the studied concentration range can be obtained by the three-parameter equations. Among all the tested equations, an excellent and perfect representation of the experimental results is obtained using the RedlichePeterson model. and the mechanism of adsorption is a hybrid unique and does not follow ideal monolayer adsorption. Kinetic study showed that sorption behavior of color, TSS, COD, BOD, and tannin and lignin onto banana peel biosorbents had a better t with the pseudo second order equation than the Lagergren rst order and second order equations. Following the pseudo second-order kinetics indicates that the biosorption process operates through chemisorption mechanism. Chemical sorption can occur by the active functional groups of biosorbents as chemical bonding agents. The addition of KOH raised the pore size and porosity of BPAC which is due to the aggressive action on the cellular structure, indicating that carbon gasication was enhanced by the addition of KOH which led to widened pores. The present work reveals that the waste banana peel is a promising material for the treatment of biologically treated POME.

248

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

Fig. 10. Surface images of (a) NBP, (b) MBP, (C1) BPAC, (C2) BPAC with the white spheres and uffy materials and (C3) BPAC with large pores.

Acknowledgment This work has been carried out during sabbatical leave granted to the author (Rae R. Mohammed) from Mosul Technical Institute during the academic year 2012, so the author thanks Ministry of Higher Education/Iraq, and Foundation of Technical Education/ Iraq, for their support. Authors would like to acknowledge and extend their heartfelt gratitude to Dr. Vasanthi Sethu who has made the completion of this research possible. We are deeply

indebted to Andrew, Y.S., Farahwahida B.M., and Filza M.F. for their technical assistance. This research has been supported by the University of Nottingham/Malaysia Campus as a research scholar fellowship. Appendix A. Supplementary data Supplementary data related to this article can be found at http:// dx.doi.org/10.1016/j.jenvman.2013.11.031

R.R. Mohammed, M.F. Chong / Journal of Environmental Management 132 (2014) 237e249

249

References
Ahmad, A.L., Chong, M.F., Bhatia, S., Ismail, S., 2006. Drinking water reclamation from palm oil mill efuent (POME) using membrane technology. Desalination 191, 35e44. Azhari, S.B., Lim, S.H., Mohd, Z.Y., NorAini, A.R., Umi kalsom, M.S., Mohd, A.H., Minato, W., Kenji, S., Yoshihito, S., 2010. Effects of palm oil mill efuent (POME) anaerobic sludge from 500 m3 of closed anaerobic methane digested tank on pressed-shredded empty fruit bunch (EFB) composting process. Afr. J. Biotechnol. 9, 2427e2436. Babel, K., Jurewicz, K., 2004. KOH activated carbon fabrics as super capacitor material. J. Phys. Chem. Solid 65, 275e280. Baig, T.H., Garcia, A.E., Tiemann, K.J., Gardea-Torresdey, J.L., 1999. Adsorption of heavy metal ions by the biomass of Solanum Elaeagnifolium (silverleaf nightshade). In: Proceedings of the 1999 Conference on Hazardous Waste Research, pp. 131e142. Chanida, R., Poonsuk, P., 2011. Decolourisation and phenol removal of anaerobic palm oil mill efuent by Phanerochaete chrysosporium ATCC 24725. In: TIChE International Conference, Thailand, pp. 1e3. Chotwattanasak, J., Puetpaiboon, U., 2011. Full scale anaerobic digester for treating palm oil mill wastewater. J. Sustain. Energy Environ. 2, 133e136. Clean Water Team, 2013. Colour of Water Fact Sheet, Fact Sheet 3.1.5.9, State Water Resources Control Board. Available from: http://www.waterboards.ca.gov/water_ issues/programs/swamp/docs/cwt/guidance/3159.pdf (accessed 20.02.13.). Cong, L., Huu, H.N., Wenshan, G., Kuo-Lun, T., 2012. Optimal conditions for preparation of banana peels, sugarcane bagasse and watermelon rind in removing copper from water. Bioresour. Technol. 119, 349e354. Febrianto, J., Kosasih, A.N., Sunarso, J., Ju, Y.H., Indraswati, N., Ismadji, S., 2009. Equilibrium and kinetic studies in adsorption of heavy metals using biosorbent: a summary of recent studies. J. Hazard. Mater. 162, 616e645. Goud, V.V., Mohanty, K., Rao, M.S., Jayakumar, N.S., 2005. Phenol removal from aqueous solutions using tamarind nut shell activated carbon: batch and column study. Chem. Eng. Technol. 28, 814e821. Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process Biochem. 34, 451e465. Hossain, M.A., Ngo, H.H., Guo, W.S., Nguyen, T.V., 2012. Palm oil fruit shells as biosorbent for copper removal from water and wastewater: experiments and sorption models. Bioresour. Technol. 113, 97e101. Hsu, L.Y., Teng, H., 2000. Inuence of different chemical reagents on the preparation of activated carbons from bituminous coal. Fuel Process. Technol. 64, 155e166. Igwe, J.C.I., Onyegbado, C.O., Abia, A.A., 2010a. Adsorption isotherm studies of BOD, TSS and colour reduction from palm oil mill efuent (POME) using boiler y ash. Ecl. Qum. 35, 195e208. So Paulo. Igwe, J.C.I., Onyegbado, C.O., Abia, A.A., 2010b. Studies on the kinetics and intraparticle diffusivities of BOD, colour and TSS reduction from palm oil mill efuent (POME) using boiler y ash. Afr. J. Environ. Sci. Technol. 4, 392e400. Jamil, R.M., Saima, Q.M., Muhammad, I.B., Muhammad, Y.K., 2008. Banana peel: a green and economical sorbent for Cr (III) removal. Pak. J. Anal. Environ. Chem. 9, 20e25. Khan, T., Azhari, M.A.B., Chaudhuri, M., 2012. Banana Peel: a low-cost adsorbent for removal of reactive dye from aqueous solution. In: Proceedings of the International Conference on Civil, Offshore and Environmental Engineering. Universiti Teknologi Petronas, Malaysia. Kutty, S.R.M., Ngatenah, S.N.I., Johan, N.A., Amat, K.A.C., 2011. Removal of Zn (II), Cu (II), chemical oxygen demand (COD) and colour from anaerobically treated

palm oil mill efuent (POME) using microwave incinerated rice husk ash (MIRHA). In: International Conference on Environment and Industrial Innovation IPCBEE, Singapore, pp. 90e94. Malaysian Palm Oil Council (MPOC), 2013. Available from: http://www.mpoc.org. my/Default.aspx (accessed 10.01.13.). Malik, P.K., 2004. Dye removal from wastewater using activated carbon developed from sawdust: adsorption equilibrium and kinetics. J. Hazard. Mater. 113, 81e88. Mohamed, M.M., 2003. Acid dye removal: comparison of surfactant-modied mesoporous FSM-16 with activated carbon derived from rice husk. J. Colloid Interface Sci. 272, 28e34. Mopoung, S., 2008. Surface image of charcoal and activated charcoal from banana peel. J. Microsc. Soc. Thail. 22, 15e19. Ngah, W.S.W., Fatinathan, S., 2006. Chitosan akes and chitosaneGLA beads for adsorption of p-nitrophenol in aqueous solution. Colloids Surfaces A Physicochem. Eng. Aspects 277, 214e222. Ooi, B.S., 2006. Treatment of Palm Oil Mill Secondary Efuent (POMSE) Using Fenton Oxidation System (MSc. Thesis). Universiti Teknologi Malaysia. Saifuddin, N., Dinara, S., 2011. Pretreatment of palm oil mill efuent (POME) using magnetic chitosan. E-J. Chem. 8, 67e78. Shakila, B.A., 2008. Tertiary Treatment of Palm Oil Mill Efuent (POME) Using Hydrogen Peroxide Photolysis Method (MSc. Thesis). Faculty of Civil Engineering, Universiti Teknologi Malaysia. Shavandi, M.A., Haddadian, Z., Ismail, M.H., Abdullah, S.N., Abidin, Z.Z., 2012. Removal of Fe (III), Mn (II) and Zn (II) from palm oil mill efuent (POME) by natural zeolite. J. Taiwan Inst. Chem. Eng. 43, 750e759. Tengku, E.M., Sultan, A.I., Hakimi, M.I., 2012. Vermiltration of palm oil mill efuent (POME). In: UMT 11th International Annual Symposium on Sustainability Science and Management, Terengganu, Malaysia, pp. 1292e1297. Vzquez, G., Gonzlez-lvarez, J., Garca, A.I., Freire, M.S., Antorrena, G., 2007. Adsorption of phenol on formaldehyde pre-treated Pinus pinaster bark: equilibrium and kinetics. Bioresour. Technol. 98, 1535e1540. Verma, V.K., Mishra, A.K., 2010. Kinetic and isotherm modelling of adsorption of dyes onto rice husk. Glob. NEST J. 12, 190e196. Verma, V.K., Mishra, A.K., 2008. Removal of dyes using low cost adsorbents. Indian J. Chem. Technol. 15, 140e145. Yeong, T.W., Abdul Wahab, M., Jamaliah, M.J., Nurina, A., 2010. Pollution control technologies for the treatment of palm oil mill efuent (POME) through end-ofpipe processes. J. Environ. Manag. 91, 1467e1490. Yong, B.J., Tiehu, Li, Li, X.W., Qilang, L., 2007. Preparation of activated carbons by microwave heating KOH activation. Appl. Surf. Sci. 254, 506e512. Yupeng, G., Shaofeng, Y., Kaifeng, Y., Jingzhe, Z., Zichen, W., Hongding, X., 2002. The preparation and mechanism studies of rice husk based porous carbon. Mater. Chem. Phys. 74, 320e323. Zahangir, M.A., Emad, S.A., Suleyman, A.M., Nassereldeen, A.K., 2009. The factors affecting the performance of activated carbon prepared from oil palm empty fruit bunches for adsorption of phenol. Chem. Eng. J 155, 191e198. Zahrim, A.Y., Rachel, F.M., Su, S.Y., Melvin, F., Chan, E.S., 2009. Decolourization of anaerobic palm oil mill efuent via activated sludge-granular activated carbon. World Appl. Sci. J. (Spec. Issue for Environ.) 5, 126e129. Zawani, Z., Luqman, C.A., Thomas, S.Y., 2009. Equilibrium, kinetics and thermodynamic studies: adsorption of Remazol Black 5 on the palm kernel shell activated carbon (PKS-AC). Eur. J. Sci. Res. 37, 63e71.

Вам также может понравиться