Vous êtes sur la page 1sur 10

Colloids and Surfaces A: Physicochem. Eng.

Aspects 490 (2016) 74–83

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Activated carbons produced by pyrolysis of waste potato peels: Cobalt


ions removal by adsorption
George Z. Kyzas, Eleni A. Deliyanni ∗ , Kostas A. Matis
Division of Chemical Technology, Department of Chemistry, Aristotle University of Thessaloniki, GR-541 24 Thessaloniki, Greece

h i g h l i g h t s g r a p h i c a l a b s t r a c t

• Removal of cobalt ions by activated


carbons.
• Activation with H3 PO4 at 400, 600,
800 ◦ C.
• SEM, BET, FTIR, DTG, Boehm titrations
as characterization techniques.
• Effect of pH, contact time, tempera-
ture on adsorption.

a r t i c l e i n f o a b s t r a c t

Article history: This study investigates the use of activated carbons after pyrolysis of waste potato peels for the removal
Received 20 July 2015 of cobalt ions from synthetic wastewaters (composed of various Co(II) concentrations in distilled water
Received in revised form 9 November 2015 which adjusted previously at different pH values) without any co-existing ions. The activation agent
Accepted 16 November 2015
was phosphoric acid, while three different carbons were prepared based on the different activation
Available online 19 November 2015
temperatures (400, 600, 800 ◦ C). Characterization techniques used for the study of surface chemistry of
carbons prepared (Boehm and potentiometric titrations, BET analysis, SEM/EDAX, DTG), while the possi-
Keywords:
ble adsorption mechanism between carbons and Co2+ was further investigated using FTIR spectroscopy.
Cobalt ions
Activated carbons
Equilibrium and kinetic experiments were carried out at the optimum pH (6). Free energy, enthalpy, and
Phosphoric acid activation agent entropy were also evaluated.
Adsorption © 2015 Elsevier B.V. All rights reserved.
Kinetics

1. Introduction in many places; hence making it necessary, particularly for devel-


oping countries, to find a cheap and available feedstock for the
The most widely used carbonaceous materials for the industrial preparation of activated carbon for use in industry, drinking water
production of activated carbons are coal, wood and coconut shell. purification and wastewater treatment. Biosorption, a biological
These types of precursors are quite expensive and often imported, method of environmental control was proposed as an alternative
to conventional waste-treatment facilities [1]. Several suitable agri-
cultural by-products (lignocellulosics) – i.e., including olive-waste
∗ Corresponding author. Fax: +30 2310 997859. cakes [2], cattle-manue compost [3], bamboo materials [4], apple
E-mail addresses: georgekyzas@gmail.com (G.Z. Kyzas), lenadj@chem.auth.gr pulp [5], potato peel [6] – have been investigated in the last years as
(E.A. Deliyanni), kamatis@chem.auth.gr (K.A. Matis).

http://dx.doi.org/10.1016/j.colsurfa.2015.11.038
0927-7757/© 2015 Elsevier B.V. All rights reserved.
G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83 75

activated carbon precursors and are still receiving renewed atten- impregnated with 125 mL of H3 PO4 (75% w/w) at room temper-
tion. ature and kept under stirring for 24 h. The impregnated biomass
Activated carbons possess large surface area, as mentioned was dried in sand bath at high temperature in order to remove
above and different surface functional groups, which include car- residual water and then oven-dried for 24 h at 100 ◦ C. A weighed
boxyl, carbonyl, phenol, quinone, lactone and other groups bound amount of the dried impregnated samples was placed in a fur-
to the edges of the graphite-like layers. Therefore, they are regarded nace and heated up to (i) 400, (ii) 600, and (iii) 800 ◦ C activation
as good adsorbents for the removal of heavy metal ions and other temperatures. The treatments were carried out with nitrogen flow
inorganic substances, as well as many organic compounds from (99.999% pure) of 30 STP cm3 /min; the latter was kept at the same
liquid and gas phases [7]. Both surface chemistry and texture of rate during heating and cooling, and at a constant heating rate of
carbons are affected by the nitrogen source and the type of oxygen 10 K/min. The treatment at the activation temperature lasted 2 h
functionalities preexisting on the surface [8,9]. Acid–base interac- and after cooling the solid residues were washed in a Soxhlet appa-
tions were found to essentially govern the ammonia removal by ratus for 24 h until neutral pH and then with ethanol. The resulted
modified activated carbons [10]. activated carbons were dried at 100 ◦ C for 24 h. In order to make
A mercury removal method using polymer-coated activated car- easily understood the abbreviations used for the activated carbons
bon was studied for possible use in water treatment [11]; in order samples, they were designated as PoP400, PoP600 and PoP800 (the
to increase the affinity of activated carbon for mercury, a sulfur-rich term “Po” corresponds to potato peels), “P” to phosphoric acid used
compound, polysulfide-rubber polymer, synthesized by condensa- for activation, and the numbers “400”, “600”, and “800” to 400, 600,
tion polymerization, was coated onto the activated carbon. The role 800 ◦ C activation temperatures, respectively.
of surface functionality on silica and carbonaceous materials for
adsorption of cadmium ion was also examined using various meso- 2.3. Characterization—instrumentation
porous silica and activated carbon [12]. Introduction of nitrogen
functional groups, such as amine, pyridinic and pyrolic onto the sur- Scanning electron microscopy (SEM) images were performed
face of adsorbents and replacing them with oxygen groups is known with electron microscope (model Zeiss Supra 55 VP, Jena, Germany)
as one of the modification methods for producing strong adsor- equipped with an energy dispersive X-ray (EDX) Oxford ISIS 300
bents toward heavy metals, like Cu(II) ion; a new procedure for micro-analytical system. The accelerating voltage was 15.00 kV,
nitrogenation of commercial activated carbon was proposed [13]. while the scanning was performed in situ on a sample powder.
An activated charcoal, produced from an agricultural waste product EDAX analysis was done at magnification 10 K and led to the maps
(Lapsi seed), was tested for heavy metal ions removal; according to of elements and elemental analysis for the investigation of Co(II)
the experimental results the kinetics of the sorption process was distribution on carbon surface.
found to be pseudo-second order [14]. Thermal stability of the samples was measured (6 mg of sample
Cobalt is a strategic metal (usually obtained from different ores), was used for each measurement) by thermogravimetric analysis
used widely in the manufacture of catalysts, alloys, steels etc. Var- (TGA) using Pyris 1 TGA thermal analyzer (PerkinElmer, Dresden,
ious separation processes were applied for cobalt ions, including Germany) with the following settings: (i) 10 K/min as heating rate,
radioactive ones [15]. Especially, adsorption as one of the most and (ii) 100 mL/min as flow rate of nitrogen atmosphere.
promising techniques [16–26] is applied for Co(II) removal [27–29]. The FTIR spectra of the samples were taken with a Nicolet 560
Activated carbon adsorption has been certainly applied; the adsor- (Thermo Fisher Scientific Inc., MA, USA) FTIR spectrometer. The
bent was either commercial [30], synthesized from a sugarcane spectra were recorded in transmission mode using potassium bro-
industry waste [31] or apricot stone [32], among other sources. mate wafers containing 0.5 wt% of carbon. Pellets made of pure
The aim of this study is to prepare activated carbons from potato potassium bromate were used as the reference sample for back-
peels and test them as adsorbents for Co(II) ions. Of course there are ground measurements. The spectra were recorded from 4000 to
works regarding the use of potato peels for adsorption applications 400 cm−1 at a resolution of 4 cm−1 and are baseline corrected.
[33–36], but none with cobalt ions as under-removal pollutant. Nitrogen isotherms were measured using AS1Win (Quan-
Moreover, the effect of activation temperature was evaluated tachrome Instruments, FL, USA) at liquid N2 temperature (77 K).
with characterization techniques (Boehm and potentiometric titra- The samples were degassed at 150 ◦ C in a vacuum system at 10−4
tions, BET, SEM/EDAX, FTIR, DTG) and the best adsorbent samples Torr before the analysis. The specific surface area (SBET ) was calcu-
were further used to adsorption experiments (pH-effect, isotherms, lated from the isotherm data using the Brunauer, Emmet and Teller
kinetics, thermodynamics). (BET) model. The micropore volumes (Vmic ) were obtained with the
accumulative pore volume using density functional theory (DFT)
method. The total pore volumes (Vtot ) were obtained from the vol-
2. Materials and methods
umes of nitrogen adsorbed at a relative pressure of 0.99 cm3 /g. The
mesopore volumes (Vmes ) were calculated by subtracting Vmic from
2.1. Chemical reagents
Vtot . The pore size distribution curves were also obtained using DFT
method.
H3 PO4 (for potato peels activation) and Co(NO3 )2 ·H2 O were
For the measurement of the surface pH of carbon samples, 0.4 g
purchazed from Sigma–Aldrich (St. Louis, MO, USA); they were of
of dry carbon sample were added to 20 mL of water, and the sus-
analytical grade (puriss. pa. ≥ 98.5%).
pension was stirred overnight to reach equilibrium. Then the pH
of the solution was measured. This method provided information
2.2. Preparation of activated carbons about the acidity or basicity of the material’s surface.
The determination of surface functional groups was based on
Waste potato peels were washed with distilled water in order the Boehm titration method [37]. Aqueous solutions of NaHCO3
to remove dust and other inorganic impurities. Then, they were (0.05 mol/L), Na2 CO3 (0.05 mol/L), NaOH (0.10 mol/L), and HCl
filtered and oven-dried for 24 h at 393 K for achieving the reduction (0.10 mol/L) were prepared. 25 mL of these solutions were added to
of moisture content. The dried biomass was ground and sieved to vials containing 0.5 g of adsorbents, let to be shaken (140 rpm) until
particles (+0.45–0.15 mm). The latter was denoted as “Po”. equilibrium (24 h), and then filtered. Five blank solution (without
The chemical activation of dry biomass was achieved with adsorbent) were also prepared. In this way, the number of the basic
phosphoric acid (H3 PO4 ). 10 g of dry potato peels precursor were sites was calculated from the amount of HCl which reacted with
76 G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83

adsorbent. The various free acidic groups were derived using the 2.4.3. Effect of initial ion concentration—isotherms
assumption that: (i) NaOH neutralized carboxyl, lactonic and phe- The effect of initial ion concentration on equilibrium was
nolic groups; (ii) Na2 CO3 neutralized carboxyl and lactonic groups, observed by mixing 1 g/L of adsorbents with 20 mL of ion solutions
and (iii) NaHCO3 neutralized only carboxyl groups. The excess of of varying initial concentrations (C0 = 10–1000 mg/L). The suspen-
base or acid was then determined by back titration using NaOH sions were shaken for 180 min (optimum contact time found) at pH
(0.10 mol/L) and HCl (0.10 mol/L) solutions [38]. 6 in water bath at 25 ◦ C (N = 140 rpm).
Potentiometric titration measurements were carried out with The experimental equilibrium data were fitted to the Lang-
a T50 automatic titrator (Mettler Toledo, Columbia, USA). 0.1 g of muir (Eq. (4)) [44] and Freundlich (Eq. (5)) [45] isotherm equations
carbon were placed in 50 mL of NaNO3 solution (0.01 mol/L) as sup- expressed by the following equations:
porting electrolyte and stirred overnight. The solution was titrated
Qm KL Ce
with NaOH (0.1 mol/L) under N2 saturation. The titrations were car- Qe = (4)
1 + KL C e
ried out over a wide range of pH. The total surface charge, Qsurf
(mmol/g), was calculated as a function of pH by the following equa- Qe = KF Ce 1/n (5)
tion [39]:
where Qm (mg/g) is the maximum amount of adsorption; KL (L/mg)
CA + CB +[H+ ]+[OH− ] is the Langmuir adsorption equilibrium constant; KF (mg1−1/n
Qsurf = (1)
W L1/n /g) is the Freundlich constant representing the adsorption
capacity; n (dimensionless) is the constant depicting the adsorption
where CA and CB (mol/L) are the acid and base concentrations,
intensity.
respectively; [H+ ] and [OH− ] are the equilibrium concentrations
The equilibrium amount in the solid phase (Qe , mg/g) was cal-
of those ions (mol/L), and W (g/L) is the solid concentration.
culated according to the following equation:
The analysis of residual concentration of Co(II) was achieved
using flame atomic absorption spectrophotometer (PerkinElmer (C0 − Ce ) V
Qe = (6)
1100 B, Shelton, USA). The absorbance was converted to concen- m
tration using calibration curve.
2.4.4. Effect of temperature—thermodynamics
To study the effect of temperature on equilibrium, completely
2.4. Adsorption experiments
same experiments as those of 2.4.3 Section, but just varying the
temperature (25, 45, and 65 ◦ C). The scope of the above was also
2.4.1. Effect of pH on adsorption
to investigate the thermodynamic behavior of the carbons-ions
The effect of pH was conducted by mixing 1 g/L of adsorbent
system. For this purpose, three main parameters have been cal-
(m = 0.02 g) with 20 mL (V) of ion solution (C0 = 200 mg/L). The pH
culated (the change of Gibbs free energy (G0 , kJ/mol), enthalpy
value, ranging between 2 and 6, was kept constant throughout
(H0 , kJ/mol) and entropy (S0 , kJ/mol K)) based on the isotherms
the adsorption process by micro-additions of HNO3 (0.01 mol/L)
resulted at 25, 45, and 65 ◦ C. The system of equations below can
or NaOH (0.01 mol/L). The pH adjustment of Co(II) solutions in the
be used for the calculation of the aforementioned thermodynamic
presence of adsorbent was realized in order to study and under-
parameters (where Cs (mg/L) is the amount adsorbed on a solid at
stand the possible adsorption mechanisms in each particular pH
equilibrium and R (8.314 J mol−1 K−1 ) is the universal gas constant)
condition. pH-effect experiments were not carried out at pH > 6,
[46]:
in order to avoid precipitation phenomena of Co(II). In this pH-
zone, the metal ions get out of the solution due to formation of Cs
Kc = (7)
colloidal precipitate of Co(OH)2 and not due to the adsorption of Ce
free Co(II) ions [40]. In general the uptake of those metals at pH > 6
G0 = −RTln (Kc ) (8)
is attributed to the formation of metal hydroxide species such as
0 0 0
soluble Co(OH)+ and/or insoluble precipitate of Co(OH)2 . The sus- G = H − TS (9)
pension was shaken for 24 h (agitation rate: N = 140 rpm) into a  
H 0 1 S 0
water bath to control the temperature at 25 ◦ C (model Grant Instru- ln (Kc ) = − + (10)
R T R
ments OLS Aqua Pro, Cambridge, UK). The optimum pH found (for
performing the equilibrium experiments) was 6.
G0 was given from Eq. (8), while H0 and S0 were given
from the slop and intercept of the chart between ln(Kc ) versus 1/T
2.4.2. Effect of contact time (Eq. (10)).
Kinetic experiments were performed by mixing 1 g/L of adsor-
bent with 20 mL of ion solution (C0 = 200 mg/L). The suspensions 3. Results and discussion
were shaken for 24 h at pH 6 (optimum pH found) in water bath
at 25 ◦ C (N = 140 rpm). Samples were collected at fixed intervals 3.1. BET and thermogravimetric analysis—surface properties
(5 min–24 h). The experimental kinetic data were fitted to two
kinetic models; pseudo-first [41] (Eq. (2)) and -second order [42] The carbonization of potato peels at 400, 600 and 800 ◦ C in
(Eq. (3)) (their selection is based on the fact that they are the two the presence of H3 PO4 resulted in the formation of the activated
most widely-used kinetic models in adsorption works [42,43]). carbons PoP400, PoP600, PoP800 in yields of 55, 51, and 49%,
respectively. The increase of carbonization temperature caused
Ct = C0 − (C0 − Ce ) (1 − e−k1 t ) (2)
a progressive decrease of the yield. Similar decrease in yield at
 1
 temperatures of 800 ◦ C was also observed for carbons prepared
Ct = C0 − (C0 − Ce ) 1 − (3) from other carbonaceous precursors, polymers [47] and fruit stones
1 + k2 t
[48,49]. This can be attributed to thermal degradation of the
where k1 , k2 (min−1 ) are the rate constants for the pseudo-first, phosphor-carbonaceous species and reduction of the phosphates
-second order kinetic equations; C0 , Ct , Ce (mg/L) are the initial, present to elemental phosphorus having as result the forma-
transient and equilibrium concentrations of Co(II) in the aqueous tion of volatile phosphorous compounds (P4 O10 (phosphorus(V)
solution, respectively. oxide) and elemental phosphorus). Elemental phosphorus was
G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83 77

Table 1 0.20
Parameters of the pore structure calculated from nitrogen adsorption isotherms. 1.0 PoP400
Adsorbent 2
SBET cm /g 3
Vmic cm /g Vmes cm /g 3 3
Vtot cm /g PoP600
0.8 Po

Weight loss derivative (%/ C)


PoP400 904.56 0.099 0.357 0.726

o
0.15 0.6
PoP600 1041.43 0.055 0.336 2.960
PoP800 0.91 0.000 0.004 0.004
0.4

0.10 0.2
0.00025
Incremental pore volume (cm /g)

0.04 0.0
0.00020 100 200 300 400 500 600 700 800 900
3

0.00015
0.05
0.03 0.00010

0.00005

0.00000 0.00
0.02 10 100
100 200 300 400 500 600 700 800 900
PoP400
o
PoP600 Temperature ( C)
-in

PoP800
om

0.01 Fig. 2. Differential thermogravimetry (DTG) curves in nitrogen.


zo

0.00
sented a sufficient amount of micro and mesopores, as well as
10 100 macropores. The measured values of Vmic and Vmeso for all carbons
Pore width (A) are presented in Table 1. On the contrary, PoP800 showed a dra-
matic decrease in the amount of pores, which are mainly mesopores
Fig. 1. Comparison of the pore size distribution (DFT) curves for the PoP400, PoP600
(inset of Fig. 1). This activation temperature (800 ◦ C) caused a col-
and PoP800 carbons which activated at different temperatures.
lapse in the texture of the activated carbon prepared. Based on the
above, it was decided to not further examine the activated carbon
also formed after the phosphoric acid activation of phosphorus- sample of PoP800.
containing phenol resins [50] or Nomex polymer fibers [51]. Similar tendency was also observed for activated carbons pro-
Volatile phosphorus compounds may be formed according the fol- duced from fruit stones activated with phosphoric acid [48] and
lowing reactions [50]: was attributed to the contraction of the material [53]. The con-
4H3 PO4 +10C → P4 +10CO+6H2 O (11) traction can be due to the catalytic activity of phosphoric acid,
causing the growth of polyaromatic clusters. As a result, a densely
4H3 PO4 +10C → P4 O10 +6H2 O (12) packed structure is formed and causes loss of porosity. Additionally,
by increasing carbonization temperature up to 800 ◦ C, polyphos-
P4 O10 +10C → P4 +10CO (13)
phates can be progressively formed that can block the pore space
A useful tool for the characterization of the porosity of acti- decreasing this way the porosity [50].
vated carbons, which is an important parameter of adsorption The surface chemistry of carbons plays an important role on
performance, is nitrogen adsorption measurement; this provides in their adsorption performance. Surface chemistry as well as the
information about the textural characteristics and the surface area changes in the chemistry after the activation are reflected on the
which is available to the adsorbate [52]. The specific surface areas DTG curves for the initial biomass sample (Po) and for the activated
and the porosity parameters exhibited by the three activated car- carbons PoP400 and PoP600 (Fig. 2). The DTG curves for the initial
bons (PoP400, PoP600, PoP800) were calculated and presented in biomass sample as well as for carbon samples showed a continuous
Table 1. weight loss at about 100 ◦ C related to moisture release.
The surface area of PoP400 (904.56 cm2 /g) is slightly smaller For the Po sample (inset of Fig. 2), the defined peak in the
than this of PoP600 (1041.43 cm2 /g), while the total pore volume is temperature range 200–400 ◦ C, represents the decomposition of
smaller about 4 times than this of PoP600 (0.726 and 2.960 cm3 /g cellulose and hemicellulose [54,55]. In this temperature range, the
for PoP400 and PoP600, respectively). The increase of temperature chemical bonds in potato peels began to break and the lightest
from 400 to 600 ◦ C causes a 15% increase of surface area, while fur- volatile compounds were released [56]. At higher temperatures
ther increase of the carbonization temperature to 800 ◦ C resulted than 400 ◦ C, a small, poorly-shaped peak, may be indicative of lignin
in a decrease of the BET surface area, may be due to a collapse of weight loss [57]. In the case of PoP400 and PoP600, the effects of
structure. carbon acidity is demonstrated as a significant weight loss between
Pore size distribution is another important aspect of acti- 150 and 600 ◦ C, where the surface functional groups decompose
vated carbons since it characterizes the structural heterogeneity [58]. An intense peak at 200 ◦ C indicates a high contribution of
of porous materials. According to IUPAC definition, porous car- strongly acidic carboxylic groups. PoP600 appears to be more acidic
bons are classified into microporous (pore size < 2 nm), mesoporous than PoP400. The weight loss at 600 ◦ C is related to the decompo-
(pore size = 2–50 nm) and macroporous activated carbons (pore sition of phenols and other basic functionalities [8]. Finally, both
size >50 nm) [52]. The pore size distributions exhibited by the three carbons showed appreciable mass loss in the range of 850–880 ◦ C.
activated carbons (PoP400, PoP600, PoP800) were calculated by the Since these carbons are phosphoric-activated, the aforementioned
density functional theory (DFT) method and presented in Fig. 1. mass loss can be attributed to (i) the evaporation of phosphorus
According to Fig. 1, the porous texture of the three carbons is compounds by thermal decomposition of the C–O–P linkage, and
different. The results confirmed that the activated carbon PoP400 (ii) phosphate/polyphosphate reduction by carbon [50,51,59]. The
is of micro-mesoporous structure. As it can be observed, the graph surface areas under all peaks increase with an increase of activation
exhibits peaks with a maximum at 1–2 and 3–5 nm. PoP600 pre- temperature.
78 G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83

Table 2
Surface pH of the carbon samples and Boehm titration results.

Functional groups

Adsorbent Carboxyl Lactonic Phenolic Basic Total acid

pH mmol/g mmol/g mmol/g mmol/g mmol/g

PoP400 2.29 0.0125 0.0115 0.1307 0.0297 0.1608


PoP600 2.85 0.0125 0.0169 0.0316 0.0104 0.0609

0.5 100
PoP400
90 PoP600
0.0 80

70

Removal (%)
-0.5
Qsurf (mmol/g)

60
PoP600
50
-1.0
40

30
-1.5
20

-2.0 10
PoP400
0
2 3 4 5 6
2 3 4 5 6 7 8 9 10 11 12
pH
pH
Fig. 4. Effect of pH on Co(II) adsorption onto PoP400 and PoP600.
Fig. 3. Potentiometric titration curves for PoP400 and PoP600.

200
The acidic and basic surface groups were estimated after Boehm PoP600
PoP400
and potentiometric titration measurements and the results are pre- 175 pseudo-1st
sented in Table 2 and Fig. 3. It is obvious that both carbons are pseudo-2nd
acidic, as was also confirmed by the surface pH measurements 150
(Table 2) with the PoP400 carbon to present more oxygenated sur- 125
Ce (mg/L)

face groups.
100
3.2. Effect of pH 75

The pH of the solution plays an important role on the adsorption 50


of any ion onto materials. It is noteworthy to say that completely
25
reverse results can be found with the change of pH. That is even
more intense when the substrate (adsorbent materials) is full 0
of functional groups with different pKa values and the target- 0 50 100 150 200 250 300 350 400 1200 1500
pollutant is a heavy metal ion. Therefore, the first factor studied t (min)
in the present work is the effect of pH on Co(II) removal. Fig. 4
showed the aforementioned pH-effect. Fig. 5. Effect of contact time on Co(II) adsorption onto PoP400 and PoP600 (fitting
to pseudo-first and -second order equations).
The Co(II) removal at strong acidic pH values was found to be
very small (14 and 24% for PoP400 and PoP600, respectively). How-
ever, with the increase of pH from 2 to 3, 4, and 5, the Co(II) removal
for the observed effect of pH on the adsorption may be the suppres-
sharply increased both for PoP400 (28, 45, 62% at pH 3, 4, and 5,
sion of the complexation. In solutions with a low pH value, some
respectively) and PoP600 (33, 59, 79% at pH 3, 4, and 5, respec-
functional groups are protonated (or difficult to dissociate) and the
tively). At pH 6 was found to be the optimum pH value, in which the
complexation between Co(II) ions and the functional groups on
Co(II) removal was the highest (PoP400, 87.5% and PoP600, 92%).
the surfaces of the activated carbons are suppressed. Earlier work-
The effect of pH can be explained taking into the mind the pH at
ers have reported a similar pattern for the influence of pH on the
zero point of charge (pHzpc) and Co(II) speciation in solution. From
adsorption of heavy metals onto activated carbons [31].
the Co(II) speciation, it is seen that at the pH range 2–6Co(II) is
present as Co2+ . The pH at which the charge of the solid surface is
zero is referred to as the pHzpc. Above pHzpc, the surface charge 3.3. Effect of contact time
of the adsorbent is negative and below the particular pH it will
be positive. The amount of adsorption above the pHzpc was max- To proceed the optimization of adsorption conditions, the next
imum due to the interaction of Co2+ with the negatively charged important parameter is the effect of contact time on adsorption.
carbon surface. At low pH, particularly below pHzpc, the positively The investigation of this parameter can save money and energy in
charged Co2+ species may repel with carbon surface carrying equal the case of industrial-scale adsorption processes. The kinetic results
charge and thereby decreases the Co(II) adsorption. Another reason are presented in Fig. 5, while the kinetic parameters calculated after
G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83 79

Table 3 tural properties of two carbons. The same equilibrium experiments


Kinetic parameters for the adsorption of Co(II) onto PoP400 and PoP600.
were done at 45 and 65 ◦ C. The adsorption behavior was not the
Adsorbent Pseudo-first order Pseudo-second order same for the adsorbents studied. PoP400 increased its Qm from
373 mg/g (25 ◦ C) to 382 and 389 mg/g to 45 and 65 ◦ C, respec-
k1 min-1 R2 k2 min-1 R2
tively (Table 4). This difference was also calculated as percentage
PoP400 0.0272 0.997 0.0479 0.955
increase; 2.4% for the step from 25 to 45 ◦ C and 1.8% for the
PoP600 0.0344 0.997 0.0607 0.952
respective of 45–65 ◦ C. On the contrary, PoP600 showed higher
temperature influence. PoP600 increased its Qm from 405 mg/g
(25 ◦ C) to 447 and 480 mg/g to 45 and 65 ◦ C, respectively. This per-
Langmuir
500 Freundlich centage increase was 10.4% for the step from 25 to 45 ◦ C and 7.4%
for the respective of 45–65 ◦ C. Comparing the two carbons, the tem-
perature influence was approximately 5–8 times higher for PoP600
400
than PoP400 (Table 4).
At this point it is important to make a brief comparison
Qe (mg/g)

300 about the adsorption capacities presented in other recent pub-


lished works. Apricot stone activated carbon showed the highest
capacity (111 mg/g) [32], which is by far lowest than the value pre-
200
sented in the present study for the two activated carbons (PoP400
PoP400 PoP600 and PoP600, 373 and 405 mg/g, respectively). Some other stud-
o
100 25 C ies showed even lower adsorption capacities ranged from 0.56 to
o
45 C 80 mg/g for various materials such as pest moss, T. reesei, hectorite
o
65 C particles, roasted date pits, hezelnut shells, attapulgite, bentonite,
0
0 100 200 300 400 500 600 700 kaolin etc [32].
Ce (mg/L) A brief thermodynamic analysis was also done based on the
change of Gibbs free energy (G0 , kJ/mol), change of enthalpy
Fig. 6. Effect of initial concentration and temperature (25, 45, 65 ◦ C) on adsorp- (H0 , kJ/mol) and entropy change (S0 , kJ/mol K). These param-
tion of Co(II) onto PoP400 and PoP600 (fitting to Langmuir and Freundlich isotherm eters (at pH 6), at selected initial Co(II) concentrations (10, 50, 100,
equations). 200, 500, 700 mg/L) and all temperatures (25, 45, 65 ◦ C), were given
in Table 5. The negative values of G0 showed the spontaneous
adsorption of Co(II) onto PoP400 and PoP600.
fitting to pseudo-first and -second order equations are shown in
The positive values of H0 indicated the endothermic nature
Table 3.
of the process. These values were reduced with the increase of
The kinetic data were fitted better to pseudo-first order equa-
Co(II) concentration. The latter could be linked with the fact that
tion for both carbons (R2 ps1 = 0.997). Moreover, it was found that
in an endothermic process, the adsorbate species (cobalt ions in
the adsorption process was faster for PoP600 (k1 = 0.0284 min−1 )
this work) had to displace more than one water molecule for their
than PoP400 (k1 = 0.0272 min−1 ), probably due to the more oxygen
adsorption and this resulted in the endothermicity of the adsorp-
content on the surface of PoP600. The end of the adsorption was
tion process; therefore H0 would be positive. The magnitude
at 180 min (for both carbons). An interesting finding was the clear
of H0 might also give an idea (but not sure) about the type of
time kinetic regions for PoP400 and PoP600. During 0–60 min, the
adsorption. Moreover, the values of S0 were found to be pos-
removal of Co(II) was sharp, while for 60–150 more gradual. After
itive, which reflected the affinity of the adsorbent towards the
180 min a kinetic plateau was observed, suggesting the end of the
adsorbate species. In addition, the positive value of S0 suggested
process (Table 3).
increased randomness at the solid/solution interface with some
structural changes in the adsorbate and adsorbent. The adsorbed
3.4. Isotherms—thermodynamics solvent molecules, which were displaced by the adsorbate species,
gained more translational entropy than was lost by the adsorbate
The effect of initial ion concentration on adsorption was eval- species/molecules, thus allowing for the prevalence of random-
uated with the isotherms of equilibrium data. Fig. 6 shows the ness in the system. The positive S0 value also corresponded to
aforementioned effect at 25, 45, and 65 ◦ C. PoP400 and PoP600 an increase in the degree of freedom of the adsorbed [60].
increased their adsorption capacity with increase of the initial The different adsorption capacity of the carbons were also obvi-
Co(II) concentration from 10 to 1000 mg/L. After fitting to Lang- ous from scanning electron microscopy (SEM) measurements and
muir and Freundlich models (better fitting to Langmuir model the cobalt distribution onto the surface of cabons (Fig. 7).
R2 = 0.982–0.998), it was found that PoP600 presented higher max- The surface morphology of both carbon samples seems to be
imum adsorption capacity (Qm ) than PoP400. as expected; Many pores and cavities without any smooth surface
In particular, PoP400 showed 373 mg/g, while PoP600 405 mg/g. area. In the case of PoP600, the larger pores than PoP400 is obvi-
This change of 8% is due to the difference textural and struc-

Table 4
Equilibrium parameters for the adsorption of Co(II) at 25, 45, 65 ◦ C onto PoP400 and PoP600.

Adsorbent Langmuir equation Freundlich equation


◦ 2
T C Qm mg/g KL R KF mg1−1/n L1/n g-1 n R2

PoP400 25 373 0.035 0.998 57.68 3.31 0.937


45 382 0.041 0.998 63.55 3.42 0.936
65 389 0.047 0.997 69.52 3.53 0.934
PoP600 25 405 0.050 0.995 72.40 3.50 0.947
45 447 0.057 0.988 82.27 3.49 0.959
65 480 0.076 0.982 95.64 3.57 0.966
80 G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83

Table 5
Thermodynamic parameters for the adsorption of Co(II) onto PoP400 and PoP600.

PoP400 PoP600

C0 T Qe Kc G0 H0 S0 Qe Kc G0 H0 S0

mg/L K mg/g kJ/mol kJ/mol kJ/mol K mg/g kJ/mol kJ/mol kJ/mol K

10 298 9.90 99.00 −11.19 26.50 0.128 9.96 249.00 −13.44


318 9.92 124.00 −12.14 9.98 499.00 −15.65 52.94 0.226
338 9.95 199.00 −13.78 9.99 999.00 −17.98
50 298 45.98 11.44 −5.94 16.38 0.072 47.48 18.84 −7.15
318 46.78 14.53 −6.74 48.46 31.47 −8.69 43.37 0.170
338 46.99 15.61 −7.15 48.78 39.98 −9.60
100 298 90.96 10.06 −5.62 11.85 0.061 93.96 15.56 −6.69
318 91.98 11.47 −6.15 95.8 22.81 −7.88 28.78 0.123
338 93 13.29 −6.73 97.99 48.75 −10.12
200 298 174.89 6.96 −4.73 10.39 0.055 184.79 12.15 −6.08
318 178.87 8.47 −5.38 187.91 15.54 −6.91 25.52 0.108
338 182.68 10.55 −6.13 191.93 23.78 −8.25
500 298 314.68 1.70 −1.29 6.70 0.027 344.79 2.22 −1.94
318 324.16 1.84 −1.54 374.57 2.99 −2.76 19.83 0.074
338 334.57 2.02 −1.83 394.4 3.73 −3.43
700 298 354.98 1.03 −0.07 4.36 0.015 384.35 1.22 −0.48
318 364.91 1.09 −0.21 424.68 1.54 −1.09 15.92 0.056
338 374.58 1.15 −0.37 454.24 1.85 −1.60

Fig. 7. SEM images of (a) PoP400 and (c) PoP600; Cobalt distribution map after Co(II) onto (b) PoP400 and (d) PoP600.

ous. Figs. 7b, 7d show the cobalt distribution onto the surface of secondary hydroxyl groups. Three more vibrational bands were
carbons. It seems that PoP600 has more in number white dots sug- recorded at: (i) 1750 cm−1 attributed to the C O stretching vibra-
gesting that the adsorption of Co(II) onto PoP600 is greater than tion, of acetyl and uronic ester groups of hemicelluloses, as well as
the respective onto PoP400. The latter can be concluded after the the ester linkage of the carboxyl group in lignin and (ii) 1514 cm−1
adsorption evaluation (see Section 3.4). and 1456 cm−1 assigned to the skeletal C C vibration of aromatic
rings. The stretching vibrations of –CH3 at 1387 cm−1 is related
3.5. Adsorption mechanism to methyl structures. The adsorption bands between 1000 and
1350 cm−1 may be assigned to C O vibrations.
In order to understand the adsorption mechanism of Co(II) onto In the case of PoP400, the main characteristic absorption peaks
carbon surface, FTIR spectra were obtained for the PoP400 and are presented at: (i) 3100–3400 cm−1 attributed to O H stretching
PoP600 carbons before and after cobalt adsorption (Fig. 8). vibration which originated from chemisorbed water and phenolic
The FTIR spectrum of the potato peel sample (Po) presented groups (band not shown in figure), (ii) 1740 cm−1 assigned to car-
a wide band at about 1100 cm−1 that indicates the presence of boxylic acid vibrational frequencies, (iii) 1615 cm−1 related to C C
G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83 81

Po Co2+ Co2+
O
Co2+
O

*
O

II)

*
o(
Co2+ Co2+

-C
00
O

P4
Po
Transmittance

00
Co2+

P4
Po
O
H

II)
o(
C
C

0-
C

0
HO

P6
O HO O
Po

00
P6
Po Co2+ Co2+

* represents an "in-plane sigma-pair"


with * being a localized pi-electron
2000 1800 1600 1400 1200 1000 800
-1
Wavenumber (cm ) Fig. 9. Possible adsorption interactions between activated carbon samples and
Co(II).
Fig. 8. FTIR spectra of the studied carbon samples before and after Co(II) adsorption.
groups as measured from Boehm titration (Table 2) is expected to
be beneficial for the binding of metal ions.
The reactions that are reported to take place during the adsorp-
stretching vibration of the aromatic ring or highly conjugated C O tion of Co(II) on activated carbon [31]:
group, (iv) 1150 cm−1 attributed to OH stretching and bending
vibrations in C OH bonds, respectively, such as phenols, and (v) 2(COH)+Co2+ (CO− )2 Co2+ +2H+ (14)
980 cm−1 that could be related to epoxy groups. − 2+
2CO +Co (CO )2 Co − 2+
(15)
In the case of PoP600, besides the peaks at 1740 cm−1 due to the
stretching vibrations of C + O moieties in aromatic carboxyl groups where C represents solid surface.
and at 1150 cm−1 attributed to OH stretching and bending vibra- The interactions may proceed through a non-covalent combi-
tions in C OH bonds, a new peak at 1050 cm−1 appeared, which nation and not a chemical covalent bonding. Co2+ ions probably
can also be assigned to the C O bonds in phenols. On the other replaced the H+ of the phenol and carboxyl groups, and electrova-
hand, some new peaks at 1380 and 1660 cm−1 appeared. It has been lently attached on the C O groups.
demonstrated that the appearance of these peak was an indication In addition, the band at 1615 cm−1 was shifted to 1630 cm−1 ,
of carboxylate group [61]. which was mainly attributed to the dispersive interactions between
Moreover, since these carbons were phosphoric acid-activated, the ␲-electrons of the electron rich region in the aromatic ring
the FTIR spectra of phosphoric acid-activated carbons presented of PoP400 and Co(II). An increased adsorption was achieved may
absorption bands at about 1150, 1060 and 990 cm−1 . Those bands be due to an ion-exchange reaction between the metal ion and
are characteristic of phosphorus and phosphor-carbonaceous com- the delocalized protonated ␲-electrons of the graphene layer (-Cp-
pounds present in these carbons [62,63]. The peak at 990 cm−1 H3 O+ ) [67]. The latter has been also noticed for the adsorption of
may be attributed to P O C aliphatic or aromatic stretching, P O Cd(II), Hg(II) and Cr(III) on activated carbons [50].
stretching in >P OOH or P OH bending or P O P asymmetric Based on the above results, it is concluded that H3 PO4 activa-
stretching in polyphosphates. The peak at 1060 cm−1 could be tion introduced oxygen functional groups, especially aliphatic ones,
due to a combination of the P+ O− bond in acid phosphate esters which provided complexation sites for cobalt. Cation exchange,
and the symmetrical vibration of the polyphosphate chain P O P, electrostatic interaction, and surface complexation were consid-
while the band at 1150 cm−1 can be assigned to the stretch- ered as major mechanisms for the adsorption. Some possible
ing vibration of O C in the P O C aromatic groups or hydrogen adsorption interactions can be illustrated in Fig. 9. The above pre-
bonded P O groups of phosphates or polyphosphates [50,64–66]. sented results suggest that activated carbon from potato peels,
After Co(II) adsorption onto PoP400 and PoP600 carbons, the activated with phosphoric acid have a considerable promise for
band at 1740 cm−1 due to the stretching vibration of C O attributed heavy metal ion removal and as a consequence the water purifi-
to carboxyl groups and the band at 980 cm−1 related to epoxy cation.
groups were no longer visible, indicating that those groups have
reacted with the adsorbate/metal species. Additionally, the band 4. Conclusions
at 1150 cm−1 due to phenolic O H bending mode was eliminated
may be due to the dissociation of H+ , showing that these groups Three samples of activated carbons (PoP400, PoP600, PoP800)
were also involved in adsorption. were prepared after pyrolysis of waste potato peels in order to
FTIR results for the samples after cobalt adsorption also remove Co(II) from synthetic wastewaters. BET analysis showed
presented a decreased vibration intensity of O H due to the disso- that the surface area of PoP400 (904.56 cm2 /g) was slightly smaller
ciation of H+ as well as an eliminated absorption intensity of C O than this of PoP600 (1041.43 cm2 /g). . The adsorption properties
groups. This can lead to a conclusion that Co(II) interacted with the of PoP400 and PoP600 were studied, because PoP800 presented
C O moieties in phenolic and carboxylic groups during the adsorp- nearly zero specific surface area. FTIR spectroscopy confirmed
tion. It has also been measured that the final pH of the solution was the proposed adsorption mechanism which is based on interac-
less than the initial pH due to the exchange of Co2+ species with tions between oxygen-containing groups of carbons and Co(II).
H+ ions of the carbon. The presence of large amounts of surface The maximum theoretical adsorption capacity (Qm ) for PoP400
82 G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83

was 373 mg/g, while PoP600 even larger (405 mg/g). With the kinetics, isotherm and thermodynamic parameters, Colloid Surf. A. 457
increase of temperature, the adsorption capacity was improved, (2014) 125–133.
[25] W. Wang, J. Zhou, G. Achari, J. Yu, W. Cai, Cr(VI) removal from aqueous
but (comparing the two carbons) the temperature influence was solutions by hydrothermal synthetic layered double hydroxides: adsorption
approximately 5–8 times higher for PoP600 than PoP400. The performance, coexisting anions and regeneration studies, Colloid Surf. A. 457
adsorption experiments were carried out at pH 6 as it was found (2014) 33–40.
[26] S. Yang, L. Li, Z. Pei, C. Li, J. Lv, J. Xie, B. Wen, S. Zhang, Adsorption kinetics,
that it was the optimum pH value. Kinetic experiments showed that isotherms and thermodynamics of Cr(III) on graphene oxide, Colloid Surf. A.
the adsorption process was faster for PoP600 (k1 = 0.0284 min−1 ) 457 (2014) 100–106.
than PoP400 (k1 = 0.0272 min−1 ). [27] M.J. Angove, B.B. Johnson, J.D. Wells, The influence of temperature on the
adsorption of cadmium(II) and cobalt(II) on kaolinite, J. Colloid Interface Sci.
204 (1998) 93–103.
References [28] E.H. Borai, M.M.E. Breky, M.S. Sayed, M.M. Abo-Aly, Synthesis, characterization
and application of titanium oxide nanocomposites for removal of radioactive
[1] K.A. Matis, A.I. Zouboulis, I.C. Hancock, Waste microbial biomass for cadmium cesium, cobalt and europium ions, J. Colloid Interface Sci. 450 (2015) 17–25.
ion removal: application of flotation for downstream separation, Bioresour. [29] J.D. Wells, B.B. Johnson, The influence of temperature on the adsorption of
Technol. 49 (1994) 253–259. cadmium(II) and cobalt(II) on goethite, J. Colloid Interface Sci. 211 (1999)
[2] R. Baccar, J. Bouzid, M. Feki, A. Montiel, Preparation of activated carbon from 281–290.
Tunisian olive-waste cakes and its application for adsorption of heavy metal [30] A.H. Sulaymon, B.A. Abid, J.A. Al-Najar, Removal of lead copper chromium and
ions, J. Hazard. Mater. 162 (2009) 1522–1529. cobalt ions onto granular activated carbon in batch and fixed-bed adsorbers,
[3] M.A.A. Zaini, R. Okayama, M. Machida, Adsorption of aqueous metal ions on Chem. Eng. J. 155 (2009) 647–653.
cattle-manure-compost based activated carbons, J. Hazard. Mater. 170 (2009) [31] K.A. Krishnan, T.S. Anirudhan, Kinetic and equilibrium modelling of cobalt(II)
1119–1124. adsorption onto bagasse pith based sulphurised activated carbon, Chem. Eng.
[4] S.-F. Lo, S.-Y. Wang, M.-J. Tsai, L.-D. Lin, Adsorption capacity and removal J. 137 (2008) 257–264.
efficiency of heavy metal ions by Moso and Ma bamboo activated carbons, [32] M. Abbas, S. Kaddour, M. Trari, Kinetic and equilibrium studies of cobalt
Chem. Eng. Res. Des. 90 (2012) 1397–1406. adsorption on apricot stone activated carbon, J. Ind. Eng. Chem. 20 (2014)
[5] T. Depci, A.R. Kul, Y. Önal, Competitive adsorption of lead and zinc from 745–751.
aqueous solution on activated carbon prepared from Van apple pulp: study in [33] E.K. Guechi, O. Hamdaoui, Evaluation of potato peel as a novel adsorbent for
single- and multi-solute systems, Chem. Eng. J. 200–202 (2012) 224–236. the removal of Cu(II) from aqueous solutions: equilibrium, kinetic, and
[6] J.C. Moreno-Piraján, L. Giraldo, Activated carbon obtained by pyrolysis of thermodynamic studies, Desalin. Water Treat. (2015).
potato peel for the removal of heavy metal copper (II) from aqueous [34] G.Z. Kyzas, E.A. Deliyanni, Modified activated carbons from potato peels as
solutions, J. Anal. Appl. Pyrolysis 90 (2011) 42–47. green environmental-friendly adsorbents for the treatment of pharmaceutical
[7] D. Angin, Utilization of activated carbon produced from fruit juice industry effluents, Chem. Eng. Res. Des. 97 (2015) 135–144.
solid waste for the adsorption of Yellow 18 from aqueous solutions, [35] Y.A. Öktem, S.G. Pozan Soylu, N. Aytan, The adsorption of methylene blue
Bioresour. Technol. 168 (2014) 259–266. from aqueous solution by using waste potato peels; equilibrium and kinetic
[8] E. Deliyanni, T.J. Bandosz, Effect of carbon surface modification with studies, J. Sci. Ind. Res. 71 (2012) 817–821.
dimethylamine on reactive adsorption of NOx, Langmuir 27 (2011) [36] Z. Zhang, X. Luo, Y. Liu, P. Zhou, G. Ma, Z. Lei, L. Lei, A low cost and highly
1837–1843. efficient adsorbent (activated carbon) prepared from waste potato residue, J.
[9] M. Seredych, D. Hulicova-Jurcakova, G.Q. Lu, T.J. Bandosz, Surface functional Taiwan Inst. Chem. Eng. 49 (2015) 206–211.
groups of carbons and the effects of their chemical character, density and [37] H.P. Boehm, Chemical identification of surface groups, Adv. Catal. 16 (1966)
accessibility to ions on electrochemical performance, Carbon 46 (2008) 179–274.
1475–1488. [38] H. Valdés, M. Sánchez-Polo, J. Rivera-Utrilla, C.A. Zaror, Effect of ozone
[10] L.M. Le Leuch, T.J. Bandosz, The role of water and surface acidity on the treatment on surface properties of activated carbon, Langmuir 18 (2002)
reactive adsorption of ammonia on modified activated carbons, Carbon 45 2111–2116.
(2007) 568–578. [39] Q. Du, Z. Sun, W. Forsling, H. Tang, Adsorption of copper at aqueous illite
[11] E.A. Kim, A.L. Seyfferth, S. Fendorf, R.G. Luthy, Immobilization of Hg(II) in surfaces, J. Colloid Interface Sci. 187 (1997) 232–242.
water with polysulfide-rubber (PSR) polymer-coated activated carbon, Water [40] A.M. Donia, A.A. Atia, K.Z. Elwakeel, Selective separation of mercury(II) using
Res. 45 (2011) 453–460. magnetic chitosan resin modified with Schiff’s base derived from thiourea
[12] M. Machida, B. Fotoohi, Y. Amamo, T. Ohba, H. Kanoh, L. Mercier, Cadmium(II) and glutaraldehyde, J. Hazard. Mater. 151 (2008) 372–379.
adsorption using functional mesoporous silica and activated carbon, J. Hazard. [41] S. Lagergren, About the theory of so-called adsorption of soluble substances,
Mater. 221–222 (2012) 220–227. Handlingar 24 (1898) 1–39.
[13] M.H. Kasnejad, A. Esfandiari, T. Kaghazchi, N. Asasian, Effect of pre-oxidation [42] Y.S. Ho, J.C.Y. Ng, G. McKay, Kinetics of pollutant sorption by biosorbents:
for introduction of nitrogen containing functional groups into the structure of Review, Sep. Purif. Methods 29 (2000) 189–232.
activated carbons and its influence on Cu (II) adsorption, J. Taiwan Inst. Chem. [43] S. Azizian, Kinetic models of sorption: a theoretical analysis, J. Colloid
Eng. 43 (2012) 736–740. Interface Sci. 276 (2004) 47–52.
[14] R.M. Shrestha, I. Varga, J. Bajtai, M. Varga, Design of surface functionalization [44] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and
of waste material originated charcoals by an optimized chemical platinum, J. Am. Chem. Soc. 40 (1918) 1361–1403.
carbonization for the purpose of heavy metal removal from industrial waste [45] H. Freundlich, Over the adsorption in solution, Z. Phys. Chem. 57 (1906)
waters, Microchem. J. 108 (2013) 224–232. 385–470.
[15] M.M. Koutlemani, P. Mavros, A.I. Zouboulis, K.A. Matis, Recovery of Co2+ ions [46] J.M. Smith, H.C. Van Ness, Introduction to Chemical Engineering
from aqueous solutions by froth flotation, Sep. Sci. Technol. 29 (1994) Thermodynamics, McGraw-Hill, New York, U.S.A, 1987.
867–886. [47] A.M. Puziy, O.I. Poddubnaya, A. Martínez-Alonso, F. Suárez-García, J.M.D.
[16] D. Gu, J.B. Fein, Adsorption of metals onto graphene oxide: surface Tascón, Synthetic carbons activated with phosphoric—acid I. Surface
complexation modeling and linear free energy relationships, Colloid Surf. A. chemistry and ion binding properties, Carbon 40 (2002) 1493–1505.
481 (2015) 319–327. [48] A.M. Puziy, O.I. Poddubnaya, A. Martínez-Alonso, A. Castro-Muñiz, F.
[17] L. Jiang, S. Xiao, J. Chen, Removal behavior and mechanism of Co(II) on the Suárez-García, J.M.D. Tascón, Oxygen and phosphorus enriched carbons from
surface of Fe-Mn binary oxide adsorbent, Colloid Surf. A. 479 (2015) 1–10. lignocellulosic material, Carbon 45 (2007) 1941–1950.
[18] J. Liu, J. Cao, H. Chen, D. Zhou, Adsorptive removal of humic acid from aqueous [49] A.M. Puziy, O.I. Poddubnaya, A. Martínez-Alonso, F. Suárez-García, J.M.D.
solution by micro- and mesoporous covalent triazine-based framework, Tascón, Surface chemistry of phosphorus-containing carbons of
Colloid Surf. A. 481 (2015) 276–282. lignocellulosic origin, Carbon 43 (2005) 2857–2868.
[19] Z. Luo, M. Gao, S. Yang, Q. Yang, Adsorption of phenols on reduced-charge [50] M. Myglovets, O.I. Poddubnaya, O. Sevastyanova, M.E. Lindström, B. Gawdzik,
montmorillonites modified by bispyridinium dibromides: mechanism, M. Sobiesiak, M.M. Tsyba, V.I. Sapsay, D.O. Klymchuk, A.M. Puziy, Preparation
kinetics and thermodynamics studies, Colloid Surf. A. 482 (2015) 222–230. of carbon adsorbents from lignosulfonate by phosphoric acid activation for
[20] B.J. Naden, L.M. Kessell, P.F. Luckham, T.F. Tadros, Adsorption of the adsorption of metal ions, Carbon 80 (2014) 771–783.
poly(hydroxystearic acid) to TiO2 nanoparticles, studied using gel permeation [51] F. Suárez-García, S. Villar-Rodil, C.G. Blanco, A. Martinez-Alonso, J.M.D.
chromatography, Colloid Surf. A. 478 (2015) 36–44. Tascón, Effect of phosphoric acid on chemical transformations during Nomex
[21] S. Tu, F. Lv, P. Hu, Z. Meng, H. Ran, Y. Zhang, Preparation of amine-modified pyrolysis, Chem. Mater. 16 (2004) 2639–2647.
silica foams and their adsorption behaviors toward TNT red water, Colloid [52] Y. Chen, S.R. Zhai, N. Liu, Y. Song, Q.D. An, X.W. Song, Dye removal of activated
Surf. A. 481 (2015) 493–499. carbons prepared from NaOH-pretreated rice husks by low-temperature
[22] Y.J. Zhang, J.L. Ou, Z.K. Duan, Z.J. Xing, Y. Wang, Adsorption of Cr(VI) on solution-processed carbonization and H3PO4 activation, Bioresour. Technol.
bamboo bark-based activated carbon in the absence and presence of humic 144 (2013) 401–409.
acid, Colloid Surf. A. 481 (2015) 108–116. [53] F. Suárez-García, A. Martínez-Alonso, J.M.D. Tascón, Pyrolysis of apple pulp:
[23] S. Dengler, G.J.T. Tiddy, L. Zahnweh, W. Kunz, Specific ion adsorption on alkyl chemical activation with phosphoric acid, J. Anal. Appl. Pyrolysis 63 (2002)
carboxylate surfactant layers, Colloid Surf. A. 457 (2014) 414–418. 283–301.
[24] P. Sharma, B.K. Saikia, M.R. Das, Removal of methyl green dye molecule from [54] Y. Gao, Q. Yue, B. Gao, Y. Sun, W. Wang, Q. Li, Y. Wang, Comparisons of porous,
aqueous system using reduced graphene oxide as an efficient adsorbent: surface chemistry and adsorption properties of carbon derived from
G.Z. Kyzas et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 490 (2016) 74–83 83

enteromorpha prolifera activated by H4P2O7 and KOH, Chem. Eng. J. 232 [61] W. Liu, J. Zhang, C. Cheng, G. Tian, C. Zhang, Ultrasonic-assisted sodium
(2013) 582–590. hypochlorite oxidation of activated carbons for enhanced removal of Co(II)
[55] D. Liu, P. Yuan, D. Tan, H. Liu, T. Wang, M. Fan, J. Zhu, H. He, Facile preparation from aqueous solutions, Chem. Eng. J. 175 (2011) 24–32.
of hierarchically porous carbon using diatomite as both template and catalyst [62] M.S. Solum, R.J. Pugmire, M. Jagtoyen, F. Derbyshire, Evolution of carbon
and methylene blue adsorption of carbon products, J. Colloid Interface Sci. structure in chemically activated wood, Carbon 33 (1995) 1247–1254.
388 (2012) 176–184. [63] M. Jagtoyen, M. Thwaites, J. Stencel, B. McEnaney, F. Derbyshire, Adsorbent
[56] K.Y. Foo, B.H. Hameed, Utilization of rice husks as a feedstock for preparation carbon synthesis from coals by phosphoric acid activation, Carbon 30 (1992)
of activated carbon by microwave induced KOH and K 2CO 3 activation, 1089–1096.
Bioresour. Technol. 102 (2011) 9814–9817. [64] D.E.C. Corbridge, Infra-red analysis of phosphorus compounds, J. Appl. Chem.
[57] S. Román, J.M. Valente Nabais, B. Ledesma, J.F. González, C. Laginhas, M.M. 6 (1956) 456–465.
Titirici, Production of low-cost adsorbents with tunable surface chemistry by [65] S. Bourbigot, M. Le Bras, R. Delobel, P. Bréant, J.m. Trémillon, Carbonization
conjunction of hydrothermal carbonization and activation processes, mechanisms resulting from intumescence-part II. Association with an
Micropor. Mesopor. Mater. 165 (2013) 127–133. ethylene terpolymer and the ammonium polyphosphate-pentaerythritol fire
[58] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. Órfão, Modification of the retardant system, Carbon 33 (1995) 283–294.
surface chemistry of activated carbons, Carbon 37 (1999) 1379–1389. [66] R. Xie, B. Qu, K. Hu, Dynamic FTIR studies of thermo-oxidation of expandable
[59] A. Castro-Muñiz, F. Suárez-García, A. Martínez-Alonso, J.M.D. Tascón, graphite-based halogen-free flame retardant LLDPE blends, Polym. Degrad.
Activated carbon fibers with a high content of surface functional groups by Stab. 72 (2001) 313–321.
phosphoric acid activation of PPTA, J. Colloid Interface Sci. 361 (2011) [67] M. Sánchez-Polo, J. Rivera-Utrilla, Adsorbent-adsorbate interactions in the
307–315. adsorption of Cd(II) and Hg(II) on ozonized activated carbons, Environ. Sci.
[60] G.Z. Kyzas, M. Kostoglou, N.K. Lazaridis, Copper and chromium(VI) removal by Technol. 36 (2002) 3850–3854.
chitosan derivatives-equilibrium and kinetic studies, Chem. Eng. J. 152 (2009)
440–448.

Vous aimerez peut-être aussi