Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 15 de 15
Filter
Add more filters










Publication year range
1.
J Oleo Sci ; 62(11): 949-53, 2013.
Article in English | MEDLINE | ID: mdl-24200943

ABSTRACT

In this study, raw wheat bran (R-WB), a type of waste biomass (WB) was treated with Pectinase PL (P-WB), and the properties (yield percentage, carboxy group surface concentration, the solution pH, and specific surface area) of R-WB and P-WB were investigated. The surface concentration of carboxy groups on R-WB (3.56 mmol/g) was greater than that of P-WB (2.11 mmol/g). In contrast, the specific surface area of P-WB (24.98 m²/g) was greater than that of R-WB (3.25 m²/g). In addition, the adsorption of cadmium and lead ions to WB was evaluated. Adsorption of the heavy-metal ions reached equilibrium within 9 h, and the experimental data was fitted to a pseudo-second-order model. More heavy-metal ions were adsorbed onto R-WB than onto P-WB. The correlation coefficient between the amount of ions adsorbed and the number of carboxy groups or pectin exceeded 0.884 and 0.975, respectively. This study indicated that wheat bran was useful for the removal of cadmium or lead ions from aqueous solutions. The adsorption mechanism of cadmium and lead ions to WB was associated with presence of carboxy group in pectin.


Subject(s)
Cadmium/isolation & purification , Carboxylic Acids/chemistry , Chemical Phenomena , Dietary Fiber , Lead/isolation & purification , Polygalacturonase/metabolism , Waste Products , Adsorption , Dietary Fiber/analysis , Food Handling , Ions , Pectins/chemistry , Polygalacturonase/chemistry , Solutions , Time Factors , Wastewater , Water , Water Purification/methods
2.
J Oleo Sci ; 62(4): 241-4, 2013.
Article in English | MEDLINE | ID: mdl-23535312

ABSTRACT

This study investigated the removal of free residual chlorine by activated carbon (AC). ACs were prepared from coal (AC1) and coconut shell (AC2). The specific surface area of AC1 was larger than that of AC2. The removal of free residual chlorine increased with elapsed time and amount of adsorbent. The removal mechanism of free residual chlorine was the dechlorination reaction between hypochlorous acid or hypochlorite ion and AC. Moreover, AC1 was useful in the removal of free residual chlorine in tap water. The optimum condition for the removal of free residual chlorine using a column is space velocity 306 1/h; liner velocity 6.1 m/h.


Subject(s)
Biotechnology/methods , Charcoal/chemistry , Chlorine/isolation & purification , Coal , Cocos/anatomy & histology , Cocos/chemistry , Hydrogen-Ion Concentration , Time Factors , Water/chemistry , Water Purification
3.
J Oleo Sci ; 62(2): 109-14, 2013.
Article in English | MEDLINE | ID: mdl-23391535

ABSTRACT

This study investigated the regeneration of waste edible oil using a food additive (calcium silicate, CAS). Waste edible oil was prepared by combined heat and aeration treatment. Moreover, the deterioration of edible oil by combined heat and aeration treatment was greater than that by heat treatment alone. The acid value (AV) and carbonyl value (CV) increased with increasing deterioration; conversely, the tocopherol concentration decreased with increasing deterioration. The specific surface area, pore volume, and mean pore diameter of the 3 CAS formulations used (CAS30, CAS60, and CAS90) were evaluated, and scanning electron microscopic images were taken. The specific surface area increased in the order of CAS30 (115.54 m(2)/g) < CAS60 (163.93 m(2)/g) < CAS90 (187.47 m(2)/g). The mean pore diameter increased in the order of CAS90 (170.59 Å) < CAS60 (211.60 Å) < CAS30 (249.70 Å). The regeneration of waste edible oil was possible with CAS treatment. The AV reduced by 15.2%, 10.8%, and 23.1% by CAS30, CAS60, and CAS90 treatment, respectively, and the CV was reduced by 35.6%, 29.8%, and 31.3% by these 3 treatments, respectively. Moreover, the concentrations of tocopherol and free fatty acids did not change with CAS treatment. The characteristics of CAS were not related to the degree of change of AV and CV. However, the adsorption mechanism of polar and non-polar compounds generated in waste edible oil by CAS was related with the presence of silica gel molecules in CAS. The findings indicated that CAS was useful for the regeneration of waste edible oil.


Subject(s)
Calcium Compounds , Fatty Acids, Nonesterified/analysis , Food Additives , Food Handling , Hot Temperature , Oils , Oleic Acid/analysis , Oxidation-Reduction , Palmitic Acid/analysis , Silicates , Tocopherols/analysis , Waste Disposal, Fluid , Linoleic Acid/analysis , Oils/chemistry , Stearic Acids/analysis , Time Factors , alpha-Linolenic Acid/analysis
4.
J Oleo Sci ; 61(10): 593-600, 2012.
Article in English | MEDLINE | ID: mdl-23018856

ABSTRACT

The volume of waste fiber has increased rapidly in recent years, and this trend is expected to continue. In this study, therefore, we attempted to convert waste fiber to carbonaceous materials by carbonization and investigated the basic properties of the resulting carbonized fibers. The results demonstrated that pores tend to form and specific surface areas change substantially, depending on the carbonization conditions. The carbonization conditions resulting in the largest specific surface areas included a temperature increase and retention times of 2 h. Carbonization temperatures resulting in the maximum values of 1000°C were 900-1000°C for wool and 1000°C for both polyester and cotton. In particular, the specific surface area of cotton after carbonization at 1000°C was 1253 m²/g, and scanning electron microscopy (SEM) micrographs showed that cotton retained its fibrous form after carbonization. Thus, it is possible to inexpensively convert waste fibers to carbonaceous material by carbonization. The results indicate that for cotton fiber in particular, the practical application of this process to the production of low-cost fibrous activated carbon would be possible, since cotton fiber retains its fibrous form under carbonization.


Subject(s)
Carbon/chemistry , Waste Products , Carbon Fiber , Microscopy, Electron, Scanning , Surface Properties , Temperature
5.
Chem Pharm Bull (Tokyo) ; 60(8): 985-8, 2012.
Article in English | MEDLINE | ID: mdl-22863701

ABSTRACT

In this study, we investigated the surface properties of granulated boehmite with vinyl acetate (G-BE20) and measured the amount of phosphate it adsorbed and the effect of contact time and solution pH on the adsorption process. The specific surface area (144.9 m(2)/g) and the number of surface hydroxyl groups (0.88 mmol/g) of G-BE20 were smaller than those of virgin boehmite (BE), which gave a specific surface area and number of surface hydroxyl groups of 297.0 m(2)/g and 1.08 mmol/g, respectively. The amount of phosphate adsorbed increased with the temperature. The isotherm model of Langmuir was used to fit experimental adsorption equilibrium data for phosphate adsorption onto G-BE20. The calculated thermodynamic parameters show the spontaneous and endothermic nature of the adsorption process. The equilibrium adsorption onto G-BE20 was reached within 16 h and the amount of phosphate adsorbed was 8.4 mg/g. The kinetic mechanism of phosphate uptake was evaluated with two different models: the Largergren pseudo first- and pseudo second-order models. The data obtained showed a better fit to the pseudo second-order model (0.991) than to the pseudo first-order model (0.967), as indicated by the r values. The rate constants for the adsorption of phosphate onto G-BE20 were calculated as 0.481 1/h and 0.029 g/mg h. The adsorption of phosphate onto G-BE20 was the maximum in the pH range 3.0-4.0.


Subject(s)
Aluminum Hydroxide/chemistry , Aluminum Oxide/chemistry , Phosphates/chemistry , Adsorption , Hydrogen-Ion Concentration , Kinetics , Microscopy, Electron, Scanning , Solutions , Thermodynamics , Water/chemistry
6.
J Oleo Sci ; 61(4): 217-25, 2012.
Article in English | MEDLINE | ID: mdl-22450123

ABSTRACT

This study investigates the activated carbon (AC) treatment and ozone oxidation of the sulfa drugs--sulfamethoxazole (SMX), sulfamonomethoxine (SMM), sulfadimidine (SDD), and sulfadimethoxine (SDM)--in aqueous solution systems. Three AC samples were prepared from Shirasagi (AC1 and AC2) and coal (AC3), and the surface functional groups, solution pH, specific surface areas, pore volumes, and morphologies of the three samples were evaluated. The specific surface areas were in the following order: AC1 (1391 m²/g) > AC2 (1053 m²/g) > AC3 (807 m²/g). The pore volume and mean pore diameter of AC3 were greater than those of AC1 and AC2. The concentration of sulfa drugs adsorbed onto the AC samples reached equilibrium within 150 h. Experimental data of the adsorption rate were fitted to a pseudo-second-order model. The amount of sulfa drugs adsorbed onto the AC samples was in the order of SDM < SMM < SDD < SMX; the mechanism of adsorption of the sulfa drugs onto the AC samples depended on the hydrophobicity of the AC surface. The adsorption isotherm data were fitted to Freundlich and Langmuir models. Ozone was generated from oxygen gas using an A-27 ozone generator, and the complete degradation of the sulfa drugs by ozone treatment at 60 mL/min was achieved within 50 min. Ozone treatment caused the structure of the sulfa drugs to decompose via ozone oxidation.


Subject(s)
Charcoal/chemistry , Sewage/chemistry , Sulfadimethoxine/chemistry , Sulfamethoxazole/chemistry , Sulfamonomethoxine/chemistry , Sulfuric Acids/chemistry , Water Purification/methods , Adsorption , Hydrogen-Ion Concentration , Models, Chemical , Oxidation-Reduction
7.
J Oleo Sci ; 61(3): 149-54, 2012.
Article in English | MEDLINE | ID: mdl-22362146

ABSTRACT

The potential utility of the widely available waste product, wool textiles, in the adsorption of heavy metals from industrial and other wastewater systems was investigated by proxy experimentation. Carbon fiber was prepared from dyed wool (DW) by calcination at different temperatures (400, 600, 800, and 1000 °C, referred to as DW400, DW600, DW800, and DW1000, respectively). The samples were analyzed in terms of scanning electron microscope images, percentage yield, specific surface area, pore volume, and the pH of an aqueous suspension of virgin dyed wool (V-DW) or the calcined DW. The adsorption of Cu(II) and Pb(II) from aqueous solutions was studied using the batch method, and the effect of contact time and co-existence of metal ions was investigated. Cu(II) and Pb(II) adsorption increased with increasing DW calcination temperature in the order V-DW < DW400 < DW600 < DW800 < DW1000. The maximum equilibrium adsorption of Cu (II) and Pb(II) achieved with DW1000 (79% and 57%, respectively) was reached within 6 h. Fitting of the adsorption isotherm data for Cu(II) and Pb(II) adsorption onto DW1000 to the Freundlich equation was consistent with monomolecular adsorption onto a heterogeneous surface. The rate-limiting step was determined to be chemical sorption by fitting the adsorption kinetics data to pseudo first-order and pseudo second-order models, given that the pseudo second-order model best fit our data. The study demonstrated that DW1000 was useful for purification of wastewater containing Cu(II) and Pb(II).


Subject(s)
Carbon/chemistry , Copper/chemistry , Lead/chemistry , Wool/chemistry , Adsorption , Animals , Carbon Fiber , Coloring Agents , Kinetics , Microscopy, Electron, Scanning , Solutions , Temperature , Time Factors
8.
Chem Pharm Bull (Tokyo) ; 59(11): 1400-2, 2011.
Article in English | MEDLINE | ID: mdl-22041078

ABSTRACT

In this study, we investigated the surface properties of raw wheat bran (R-WB) and wheat bran treated with Pectinase PL (P-WB) to evaluating its efficacy for removal of cadmium from waste water. The concentration of cadmium ions adsorbed by them was evaluated. The concentration of carboxyl groups of R-WB (3.56 mmol/g) was greater than that of P-WB (2.11 mmol/g), which indicated that the pectin of R-WB was broken down, resulting in a decrease in the concentration of carboxyl groups due to the enzyme treatment. From the scanning electron microscope (SEM) images of P-WB, the pores were newly generated with enzyme treatment. The concentration of cadmium ions adsorbed onto R-WB was greater than that of cadmium ions adsorbed onto P-WB. These results show that the adsorption mechanism of cadmium ions onto R-WB and P-WB depends on the carboxyl groups of pectin.


Subject(s)
Cadmium/chemistry , Dietary Fiber/metabolism , Polygalacturonase/metabolism , Water Pollutants, Chemical/chemistry , Adsorption , Dietary Fiber/analysis , Ions/chemistry , Porosity
9.
J Oleo Sci ; 60(12): 609-11, 2011.
Article in English | MEDLINE | ID: mdl-22123241

ABSTRACT

Estrogens (estron: E1; estradiol: E2; estriol: E3) are major water pollutants. For the removal of estrogens, activated carbon (AC) and ozone treatments were employed, and the chemical oxygen demand (COD) and pH of a single solution and multiple solution systems were investigated. The removal of estrogens increased with the amount of AC. The percentage of removed estrogens by AC in the single solution system was greater than that in the ternary solution system. The estrogens were completely removed using ozone, which also reduced COD and pH. The removal of estrogens increased the water quality and decreased the amount of organic carbon.


Subject(s)
Charcoal , Endocrine Disruptors/isolation & purification , Estrogens/isolation & purification , Ozone , Water Pollutants, Chemical/isolation & purification , Water Purification/methods , Biological Oxygen Demand Analysis , Hydrogen-Ion Concentration , Solutions , Water Quality
10.
J Oleo Sci ; 60(12): 619-25, 2011.
Article in English | MEDLINE | ID: mdl-22123243

ABSTRACT

Carbonaceous material for the removal of fluoride ions from water was prepared from coffee grounds (CGs) by calcination and subsequent HCl treatment. The characteristics of the CGs, including the surface area, mean pore diameter, pore volume, and surface functional groups were determined, and the morphological characteristics were evaluated using scanning electron microscopy. The adsorption isotherms, saturated amount of fluoride ions adsorbed, and the effect of contact time and temperature on the adsorption of fluoride ions were investigated for a sample of tap water. The specific surface area of CG calcined at 600° (CG600) was larger than that of CGs calcined at 400, 800, and 1000°. Phenolic, lactonic, and carboxyl groups were detected on the CG600 surface. The adsorption capacity of the carbonized CGs for fluoride was ranked in the order CG400 < CG1000 < CG800 < CG600 (where the numeral indicates the carbonization temperature), whereas virgin CG and CG600-NAT (not treated with hydrochloric acid solution) did not exhibit any adsorption ability for fluoride ions. The amount of fluoride ions adsorbed onto CG600 increased with increasing temperature and was consistent with chemical adsorption. The mechanism of adsorption of fluoride ions onto CG600 proceeded via ion exchange with chloride ions (1:1) present on the surface of CG600. The adsorption isotherms were fitted to the Freundlich and Langmuir equations. Moreover, CG600 showed an acceptable adsorption capacity for fluoride ions present in tap water.


Subject(s)
Coffee/chemistry , Fluorides/isolation & purification , Water Purification/methods , Water Supply/analysis , Water/chemistry , Adsorption , Fluorides/analysis , Hydrochloric Acid , Incineration , Ion Exchange , Ions , Temperature , Water/analysis
11.
J Oleo Sci ; 60(7): 363-8, 2011.
Article in English | MEDLINE | ID: mdl-21701100

ABSTRACT

In this study, the amount of copper or cadmium adsorbed using waste biomass (i.e., coffee grounds (CG) and rice bran (RB)) was investigated. The amount of crude protein in defatted CG (D-CG) or RB (D-RB) was greater than that in CG or RB, respectively. The amount of copper or cadmium adsorbed using CG was greater than that using RB. Additionally, the amount of copper or cadmium adsorbed was not affected by the presence of fat in CG. Adsorption data was fitted to the Freundlich equation, and the correlation coefficients were in the range of 0.794-0.991. The main adsorption mechanism was thought to be monolayer adsorption onto the surface of the waste biomass. The adsorption rate data was fitted to the pseudo-second-order model, and the correlation coefficient average was in the range of 0.891-0.945. This result showed that the rate-limiting step may be chemisorption. Moreover, the amount of copper or cadmium desorbed from CG or RB using 0.01 mol/L or 1.00 mol/L HNO(3) was investigated. Desorption with 0.01 mol/L HNO(3) resulted in the recovery of 86-97% of the copper and cadmium, indicating that copper or cadmium that was adsorbed using waste biomass was recoverable.


Subject(s)
Cadmium/isolation & purification , Copper/isolation & purification , Waste Disposal, Fluid/methods , Water Pollutants, Chemical/isolation & purification , Adsorption , Biomass , Cadmium/chemistry , Cadmium/metabolism , Coffee/chemistry , Coffee/metabolism , Copper/chemistry , Copper/metabolism , Kinetics , Nitric Acid/chemistry , Oryza/chemistry , Oryza/metabolism , Water Pollutants, Chemical/chemistry , Water Pollutants, Chemical/metabolism
12.
J Oleo Sci ; 60(3): 133-8, 2011.
Article in English | MEDLINE | ID: mdl-21343661

ABSTRACT

In the present study, the phosphate adsorption abilities of granular materials, gibbsite (GB) and cerium hydroxide (CE), granulated with ethyl cellulose and ethanol, were investigated. For suitable granular conditions using GB or CE, the percentage of binder and amount of solvent were 10% and 3.25 mL/5 g for GB and 7.5% and 2.50 mL/5 g for CE. The amount of phosphate adsorbed by granular GB or CE was similar to that adsorbed by powdered GB or CE. The adsorption isotherms of phosphate were also similar for the granular and powdered materials. The results of a column experiment showed that when GB was used, the suitable concentration of sodium hydroxide as the desorption solution was 2 mmol/L. In these experiments, the amount of phosphate adsorbed was 31.4 mg/g, the amount of phosphate desorbed was 25.3 mg/g, and the recovery percentage was 80.5%. In contrast, when CE was used, the suitable concentration of sodium hydroxide was 500 mmol/L; the amount of phosphate adsorbed was 77.7 mg/g, the amount of phosphate desorbed was 57.8 mg/g, and the recovery percentage was 74.4%. Results also indicate that granular GB and CE could be used at least five times in a column without loss of absorption or desorption abilities. Moreover, granulation of GB and CE is possible using ethyl cellulose and ethanol.


Subject(s)
Aluminum Hydroxide , Cerium , Hydroxides , Phosphates/isolation & purification , Adsorption , Cellulose/analogs & derivatives , Drug Compounding , Ethanol , Powders , Sodium Hydroxide , Solutions , Water Pollutants, Chemical/isolation & purification
13.
J Hazard Mater ; 181(1-3): 574-9, 2010 Sep 15.
Article in English | MEDLINE | ID: mdl-20605321

ABSTRACT

The specific surface area and X-ray diffraction patterns for an aluminum hydroxide gel (AHG) calcined at 300-1150 degrees C, the number of surface hydroxyl groups in the AHG, and the adsorption isotherms of phosphate on AHG were measured in order to develop a phosphate recovery agent. AHG was transformed into gamma- and alpha-alumina by the calcinations treatment. The amount of phosphate adsorbed onto AHG increases at calcining temperatures of 300-700 degrees C and decreases above a calcining temperature of 800 degrees C. It was found that AHG selectively adsorbs phosphate ions, but not other anions, and shows the highest adsorption capacity at pH 4-6. Further, the alkali resistance of AHG increased with calcination, and more than 80% of the phosphate adsorbed with an NaOH aqueous solution underwent desorption. The addition of colloidal alumina and colloidal silica resulted in the formation of granules of 500-840 microm size. The amount of phosphate adsorbed onto AHG after granulation was similar to that before granulation. Thus, the phosphate absorption capacity of AHG did not decrease after granulation suggesting that AHG can be used as a phosphate adsorbent.


Subject(s)
Aluminum Hydroxide/chemistry , Phosphates/isolation & purification , Adsorption , Hydrogen-Ion Concentration , Particle Size , Temperature , X-Ray Diffraction
14.
J Oleo Sci ; 58(10): 529-35, 2009.
Article in English | MEDLINE | ID: mdl-19745580

ABSTRACT

Bone char (BC) was prepared by carbonizing four types of animal biomass, and the adsorption of fluoride ions and elution of phosphate ions were investigated. It was found that the BC yield decreased as carbonization temperature increased, and that carbonization temperature had no significant effect on surface pH, base or acid consumptions. Fluoride ion adsorption was increased in BC produced at a low carbonization temperature. The adsorption mechanism of fluoride ion on BCs might be monolayer adsorption. BC can potentially be used to remove fluoride ions in drinking water. However, it was found that phosphate ions from BC are eluted due to adsorption of fluoride ions, and that ingestion of large amounts of phosphate ions inhibits reabsorption of calcium in the human body. Thus there is a need to study the elution behavior of phosphate ions. The adsorption mechanisms of fluoride ions onto BC would be a physical adsorption onto BC and phosphate ion in BC is exchanged to fluoride ion.


Subject(s)
Biomass , Bone and Bones/chemistry , Fluorides/chemistry , Fluorides/isolation & purification , Adsorption , Animals , Hydrogen-Ion Concentration , Phosphates/chemistry , Phosphates/isolation & purification , Surface Properties , Temperature
15.
J Oleo Sci ; 58(5): 261-6, 2009.
Article in English | MEDLINE | ID: mdl-19367082

ABSTRACT

The present study investigates (1) ozone treatment, (2) adsorption treatment using activated carbon treatment, and (3) ozone treatment with activated carbon for their efficacy in removing 17beta-estradiol (E2) present in an aqueous solution. Both ozone and activated carbon treatments for 20 min were effective in removing E2 (initial concentration, 100 mg/L). However, both treatments have been used for two processes, and the disposal time with these treatments is more than that of another treatment. In this study, ozone treatment with activated carbon was investigated with regard to the percentage of E2 removal, pH, and chemical oxygen demand (COD). The physical and chemical characteristics of the activated carbon surface were modified due to the ozone treatment: the surface was oxidized by ozone. The surface of activated carbon exhibited polarity groups and became porous after ozone treatment. The amount of E2 adsorbed onto the ozone-treated activated carbon surface was similar to that adsorbed onto the untreated activated carbon surface. The percentage of E2 removal by ozone treatment with activated carbon was greater than that by ozone treatment alone. Moreover, the rate of E2 removal by ozone treatment with activated carbon was higher than that by ozone or activated carbon treatments alone. Ozone with activated carbon treatment may be used for the removal of E2. However, this ozone treatment for the degradation of E2 may have a greater COD than that in the case of activated carbon treatment alone.


Subject(s)
Carbon/chemistry , Estradiol/chemistry , Ozone/chemistry , Adsorption , Hydrogen-Ion Concentration , Oxidation-Reduction , Oxygen/chemistry , Porosity , Solutions , Surface Properties , Water/chemistry
SELECTION OF CITATIONS
SEARCH DETAIL
...