Anda di halaman 1dari 9

Powder Technology 245 (2013) 208–216

Contents lists available at SciVerse ScienceDirect

Powder Technology
journal homepage: www.elsevier.com/locate/powtec

Mechanism of stearic acid adsorption to calcite


Slavica R. Mihajlović a,⁎, Dušica R. Vučinić b, Živko T. Sekulić a, Sonja Z. Milićević a, Božo M. Kolonja b
a
Institute for Technology of Nuclear and Other Mineral Raw Materials, Franchet d'Esperey 86 Street, 11000 Belgrade, Serbia
b
University of Belgrade, Faculty of Mining and Geology, Djusina Street 7, 11000 Belgrade, Serbia

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

Article history: This paper presents the interpretation of binding mechanism of stearic acid for calcite surface in two modi-
Received 14 November 2012 fication methods.
Received in revised form 16 April 2013 In the “dry” method, a surface dissociation of stearic acid is assumed where H + ion goes to a surface carbon-
Accepted 20 April 2013
ate ion and stearic ion is chemisorbed on primary surface center of −Ca + ion which is only available for
Available online 2 May 2013
chemisorption. The structure of adsorbed layer indicates chemisorption of stearate, but due to steric effects,
Keywords:
and the oblique or gauche conformation of hydrocarbon chains. A part of surface − Ca+ centers can be
Calcite blocked, which is explained by the result that physical adsorption dominates over 1.5%.
Stearic acid In the “wet” method, by adding stearic acid, at concentration above the critical concentration of micelle forma-
Adsorption tion in base solution, micelles are formed, and on the other side free stearic acid molecules or molecules from de-
“Dry method” veloped micelle dissociate. The resulting stearic ions can be chemisorbed on primary centers of −Ca+ ions or
“Wet method” participate in ion exchange with OH− ions from secondary surface centers. With increasing adsorption density,
the adsorbed ions and molecules, due to interactions of hydrocarbon chains, and thickness of double electrical
layer, achieve a vertical orientation and keep the trans-conformation of hydrocarbon chains.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction adsorption places and the number of available adsorption places, so


the value is always between 0 and 1. This can be presented numeri-
The ability of limestone to be used as a filler in the polymer indus- cally by Eq. (1) [10]:
try can be significantly improved by modifying the calcite surface, as
the basic mineral in limestone, with surface-active organic modifiers V
θ¼ ð1Þ
[1–5]. V∞
When aiming to obtain a hydrophobic surface of calcite in the
modification method, monocarboxylic acids with aliphatic hydrocar- where:
bon chain (also known as fatty acids and their salts) are most fre-
quently used as the surfactants. By adsorption of surfactants on V volume of adsorbate (gas) that is adsorbed
calcite, its hydrophilic surface becomes hydrophobic, and therefore V∞ volume of adsorbate (gas) that completely covers the sur-
compatible with the polymer molecules, which further improves the face in a monomolecular layer.
mechanical properties of final product [6–9].
In many cases, the adsorption is of a mixed nature (i.e. partly Furthermore, degree of surface coverage can be presented by
chemical and partly physical) [10–12]. Usually, it occurs by formation Eq. (2):
of multilayers, wherein the first layer is built on the basis of strong
′ x
chemical forces, with the following layers physically adsorbed over θ¼k ð2Þ
m
the chemisorbed layer.
In order to better define the adsorption process, mathematical ex- where: x = mass of adsorbate; m = mass of adsorbent; and k′ = co-
pressions are used that are mainly derived from the example of gas efficient of proportionality.
adsorption on a solid adsorbent. A degree of surface coverage (θ)
can be used for quantitatively defining the adsorption process. A de- ′ xmonolayer  
k ¼ 1= xmonolayer ¼ mass of adsorbate in a monolayer ð3Þ
gree of surface coverage is the ratio of the number of occupied m

⁎ Corresponding author at: Franchet d Esperey 86, 11000 Belgrade, Serbia. Tel.:
Some studies have shown that each densely packed molecule of
+381 11 3691722; fax: +381 11 3691583. fatty acids occupies 0.16 nm 2 to 0.25 nm 2 of surface minerals. Rezaei
E-mail address: s.mihajlovic@itnms.ac.rs (S.R. Mihajlović). et al. [13] found that during the “wet method” of calcite modification

0032-5910/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.powtec.2013.04.041
S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216 209

with stearic acid, the monolayer adsorption is achieved, where each 2.2. Methods
molecule of stearic acid occupies 0.165 nm 2. In the case of the “dry
method” of modification, one molecule of stearic acid in the statistical The “wet method” of calcite modification with stearic acid, as well as
monolayer, occupies a calcite surface area of 0.280 nm 2. the characterization of the obtained products has been described in a
Hence, it was possible to calculate – according to Eq. (4) – the sur- previously published paper [14]. To modify calcite with the “dry
face of calcite that was available for adsorption of one molecule of method”, six samples were prepared per 200 g of calcite, with each sam-
stearic acid [14], under set experimental conditions: ple receiving different levels of stearic acid mass. Mass of stearic acid
added to the six samples, calculated for each 100 g of calcite, included
0.5 g; 1 g; 1.5 g; 2 g; 3 g and 4 g. Constant mass of calcite (100 g) was
SA used due to the constant interaction of calcite with the surfactants.
σ¼ ð4Þ
Γ⋅NA According to the stearic acid utilized in the mixture (calcite + stearic
acid), the acid concentration in the six samples was: 0.4975%; 0.9901%;
1.4778%; 1.9608%; 2.9126% and 3.8462% respectively. Concentration in
where:
the industrial conditions is also expressed in phr units (parts per hun-
dred resin), so that the concentration of stearic acid in relation to calcite
σ surface per molecule, m 2
will be: 0.5; 1; 1.5; 2; 3 and 4 phr (phr-mass of stearic acid (g) per 100 g
SA specific surface of calcite, m 2/g
of limestone).
Γ adsorbed amount of stearic acid, mol/g
The experiments using the “dry method” of calcite surface modifica-
NA Avogadro's number, 6.022 × 10 23 molecules/mol.
tion were carried out in a vibrating mill with ring working elements,
model “MN 954/3” (manufacturer “KHD Humboldt Wedag” — Germany).
The authors previously published the results of the surface modi- The volume of a mill container carrying a sample for modification is
fication of calcite using various concentrations of stearic acid using 2 dm3. The device operates discontinuously in an open-air environment.
the “wet process” [14]. This paper describes the method of “dry mod- Modification time is 7 min at 70 °C. The obtained calcite samples, modi-
ification” of calcite with stearic acid as well as the characterization of fied by the “dry method”, are denoted as follows: CD-0.5, CD-1, CD-1.5,
modified samples. The mechanism of stearic acid binding to the sur- CD-2, CD-3 and CD-4 (where C = calcite, D = “dry method”, ad num-
face of calcite – based on the overall results of modification for both ber = concentration of stearic acid). Thermal analyses were carried out
methods – is presented in this paper. on the obtained products using the same device as for “wet method” [14].
The thermogravimetric/differential thermal (TG/DT) analysis was
carried out in air atmosphere in the temperature range of 20–1000 °C,
2. Experimental at a heating rate of 10 °C/min using a Netzsch STA-409 EP analyzer.
The effect of surface modification is evaluated by the floating test
2.1. Materials and reagents reported by Sheng et al. [15], which represents the ratio of the floated
product to the overall weight of the sample after they are mixed in
Limestone from the deposit “Venčac”-Aranđelovac (Serbia) was used water and stirred vigorously. The ratio is called active ratio.
in the experimental study as the starting sample: upper grain-size limit
of the sample — 10.4 μm; mean grain diameter (d50) — 6.2 μm; specific
surface area (S) — 4.8 m2/g; degree of whiteness — 93.1%; and moisture 3. Results and discussion
content — 0.02%.
The results of chemical analysis have shown that the dominant ox- Thermal analysis of the initial calcite sample, as well as calcite sam-
ides are CaO (54, 77%) and MgO (0.79%), while the oxides Al2O3, ples modified by “wet” and “dry methods”, were carried out in order to
Fe2O3, Na2O and K2O are present in trace amounts. Based on the content monitor the changes that have occurred in the procedure of calcite
of CaO and assuming that all MgO is associated with dolomite, it was modification, i.e. stearic acid adsorption on its surface.
calculated that the sample has 95.84% of CaCO3. The surface-active mat-
ter, used in the experiments of calcite modification is technical grade 3.1. Differential thermal analysis of modified calcite samples, obtained by
stearic acid, produced by the company “Fluka” — Switzerland. The char- the “wet method”
acteristics of stearic acid used for modification of calcite surface are
presented in Table 1 [14]. The authors have reported in detail the thermal analyses of calcite
samples obtained by the “wet modification method”, in a previously
published paper [14]. Only a part of these results will be presented
Table 1
Characteristics of the stearic acid. as an introduction to explain the mechanism of stearic acid binding
to the surface of calcite.
Name Stearic acid
TG diagrams and weight losses in the various temperature ranges
Molecular formula CH3(CH2)16COOH for the starting sample of calcite, and calcite samples modified with
Molecular weight, g/mol 284.47 stearic acid by the “wet method” are presented in Fig. 1, and DTA
Density, g/ml 0.847
Dissociation constant, pKa 5.7
curves in Fig. 2.
Solubility in water, g/100 ml 0.034 (25 °C) The main exothermal maximums (>310 °C) on DTA curves of cal-
0.1 (37 °C) cite, modified by the “wet method” (Fig. 2) are attributed to the chem-
Solubility in alcohol, g/100 ml 2.5 (cold) ical adsorbed organic component, stearic ion, CH3(CH2)16COO −, and
Solubility in ether, CHCl3, CCl4, CS2 Very soluble
(St−), which is bound to the positive surface centers in a double electric
Thermal analysis 0
0,2

0,0 layer. Since the chemical adsorption provides the strongest bond be-
-20
-0,2
tween the adsorbates and adsorbents, this exothermic maximum at
d(wt)/dT (%/oC)

-40
DTA (µV)

-60
-0,4
the highest temperature is therefore attributed to the oxidation of the
-0,6

-80
-0,8
chemically bound organic component.
-100
exo

o
341 C
-1,0
o
319 C
In the “wet modification method”, in addition to the primary surface
-120
100 200 300
o
400
-1,2
100 200 300 400 ions of calcium (−Ca +), the secondary active centers for chemisorption
Temperature ( C) o
Temperature ( C)
can be formed due to the solubility of calcite on the surface. Ion
210 S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216

−Ca+ −CaOH

+ H2O ð1Þ

−CO3 −CO3H

The ion exchange reaction is then possible between stearate ions


from solution St −, and OH− from the surface center of \CaOH, and
the presence of bicarbonate ion \CO3H on the surface. Also, CaOH+
ions from the solution can be adsorbed on the primary surface centers
of calcite \CO3−, forming the new surface centers \CO3CaOH, on
which the given reaction of ion exchange can be developed.
Therefore, it can be generally stated that in an aquatic environment,
the heterogeneous calcite surface is formed with different surface cen-
ters, which create a larger number of centers for chemisorption of the
number of primary active centers −Ca+ present on calcite. Structure
of the double electrical layer formed on the border of calcite/aquatic me-
dium affects the interaction between the minerals and surfactants — in
this case stearic acid.
The position of other exothermic maximums at lower temperatures
indicates the additional, weaker interactions between the chemical spe-
Fig. 1. TG curves and weight loss for samples modified by “wet” method [13]. cies of stearic acid (ions, monomers, acid dimers) with the chemical
species present in the solution, resulting in the dissolution of calcite
on the surface of minerals. As the chemical adsorption is monolayer,
and due to the strong chemical bond between the adsorbate and surface
exchange with stearic ions can also be formed due to the adsorption of center, the exothermic peak at T > 310 °C is attributed to it. The peaks
some ions. at lower temperatures can only be attributed to the physically adsorbed
It should be noted that the surface charge of calcite depends pri- molecules of acid as well as the surface precipitated calcium stearate,
marily on the redistribution of potential determining ions (Ca 2+ CaSt2, formed in the electric layer of minerals. Based on this result, the
and CO32−) between the surface and the solution. However, H + and peaks on DTA curves of samples at 284 °C and 217 °C are attributed to
OH − ions, as the potential determining ions of the secondary type, the molecules of stearic acid (and surface precipitated calcium stea-
i.e. acid/base reaction, particularly affect the concentration of some rate), physically adsorbed in the double electric layer of calcite by vari-
ionic and molecular species in the solution and on the surface. ous energies. Pure stearic acid has an exothermic peak at 341 °C, and
As a result, the secondary surface centers may appear on the pri- the stoichiometric compound CaSt2 has a peak on DTG at about
mary surface centers by protonization reaction: 460 °C [4,13]. This difference can be explained by different entropies
of pure compounds and in the adsorption layer of minerals.
The intensity increase of exothermic peaks indicates the presence
of increasing amounts of adsorbed component, whether chemically
adsorbed stearic ion or physically adsorbed species, which can form
multiple layers.
Weight loss in the second temperature range (200 °C to 400 °C,
Fig. 1) may be also used to calculate the statistical coverage of min-
erals [9,13,14]. The mineral surface area per molecule of stearic acid
was calculated according to Eq. (4). The results are presented in
Table 2.
Calculation of adsorbed organic component was done with an es-
timate that the weight loss in the temperature range 200 °C to 400 °C
(Fig. 1), derives only from the oxidation of the molecules of stearic
acid. If we calculate the amount of added stearic acid using the molec-
ular weight, the expected adsorbed amount with 1% of stearic acid is

Table 2
The adsorbed amount per unit mass and surface area of calcite, surface area per mole-
cule of stearic acid and coverage degree θ in samples modified by the “wet method”.

Sample Adsorbed Adsorbed Adsorbed Surface Degree of


amounta, amountb, amount, area per coverage, θ
μmol/g μmol/g μmol/m2 molecule, nm2

CS-0.5 17.6 14.76 3.075 0.540 0.30


CS-1 35.1 33.40 6.957 0.239 0.69
CS-1.5 52.7 50.97 10.619 0.156 1.05
CS-2 70.2 68.20 14.208 0.117 1.41
CS-3 105.3 84.67 17.640 0.094 1.75
CS-4 140.4 110.73 23.069 0.072 2.29
a
Adsorbed amount calculated mathematically based on the added amount and the
molecular weight Γ = [m/M]/100, where m is the amount of the added stearic acid
Fig. 2. Comparative DTA curves of starting and calcite samples modified by “wet” (0.5, 1, 2, 3 and 4 g) and M is the molecular weight of the stearic acid (284.5 g/mol).
b
method. Adsorbed amount calculated from the weight loss.
S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216 211

35.1 μmol/g. From Table 2 we see that the amount calculated from the
weight loss is 33.40 μmol/g. This indicates that, almost complete
stearic acid has been adsorbed and good overlapping of these data
confirms that we can use these weights for the calculation of the sur-
face area per molecule and degree of coverage. When the concentra-
tion of the stearic acid is >2% the difference between the results of
the adsorbed amount calculated from the weight loss and theoretical-
ly using the molecular weight, rises. This suggests that increasing
amount of acid is physically adsorbed on the surface of calcite. The
data in the second column is calculated according to the specific sur-
face area of calcite (S = 4.8 m 2/g).
The results presented in Table 2 for calcite modified by the “wet
method” show that the calculated surface area per molecule of stearic
acid decreases with the increasing of its concentration, used in the
modification from 0.540 nm 2 for sample CS-0.5 to 0.072 nm 2 for
sample CS-4. Literature has shown that when the concentration of
surface active matter (SAM) is smaller or equivalent to when a mono-
layer of organic phase is formed on the surface, or at low adsorption
density, a distribution of hydrocarbon chains depends on the concen-
tration of organic components used for modification [13,16,17]. At
low adsorption densities, SAM is bound to the surface of minerals
by electrostatic interactions and is oriented in relation to it either hor- Fig. 3. TG curves and weight loss for samples modified by “dry” method.
izontally or at an angle by a hydrocarbon chain (“tail”) (weaker ar-
rangement of adsorbed component) and, thus, it occupies the larger acid molecules that are strongly bound in the adsorption layer [19].
surface area. At higher adsorption density of SAM, the interaction be- The observed peaks in the temperature range of 400 °C to 600 °C on
tween the hydrocarbons occurs and they orientate vertically in rela- DTA curves of modified samples can be attributed to the beginning of
tion to the surface of minerals (greater arrangement, i.e. denser the transformation phase of calcite and other minerals, which are pres-
packaging of SAM) and the occupied surface area decreases. ent in trace amounts in limestone. Hence, this temperature range is not
Based on the data in Table 2, the calculated value of calcite surface further considered.
area per molecule of stearic acid of 0.54 nm2 at the lowest concentra- TG diagrams and weight loss in separate temperature ranges for
tion of stearic acid of 0.5% indicates that the alkyl chains of adsorbed or- the starting sample of calcite, as well as calcite samples modified
ganic component are most likely arranged horizontally to the surface of with stearic acid and by using the “dry method”, are presented in
calcite. With increasing concentration of stearic acid (1% and 1.5%), the Fig. 3.
adsorption density or degree of coverage increases, indicating the redis- In the first temperature range from 20 °C to 200 °C the weight loss
tribution and increasing of the vertical orientation of alkyl chains in re- can be attributed to the dehydration of minerals, i.e. loss of physisorbed
lation to the calcite surface. Using the “wet method” (Table 2), we can water as well as the weakly bound or free organic components. In the
see that the addition of stearic acid results in a gradual coverage of the
calcite surface. Only when the concentration of stearic acid exceeds 3%
the second layer is fully formed. This indicates that the stearic acid dis-
tributes evenly using this method of modification.
These results are also in agreement with the results of Wright and
Pratt, who have established that the surface area per high fatty acid
molecule from 0.22 nm 2 to 0.26 nm 2 of minerals corresponds to the
vertical orientation of alkyl chains. Furthermore, surface area above
0.515 nm 2 corresponds to the horizontal orientation [18] as well as
the obtained results of thermal DTA analysis [14].
The statistical monolayer (0.165 nm2/molecule or 10.064 μmol/m 2)
is theoretically achieved at concentrations of less than 1.5 phr stearic
acid (calculated value 1.37 phr ≈ 1.36%). The degree of coverage θ, at
concentration of 1.5 phr is 1.05, which means that statistically at that
acid concentration the second layer begins to form in addition to a
monolayer. This is in a good correlation with DTA results where the exo-
thermic peak of chemically adsorbed organic compound (monolayer
adsorption) dominates on a curve of CS-1.5 sample [14].

3.2. Differential thermal analysis of modified calcite samples, obtained by


the “dry method”

On TG and DTA curves of calcite, modified with stearic acid, three


temperature ranges are observed (Figs. 3 and 4). In the first temper-
ature range of 20 °C to 200 °C, minerals are dehydrated, i.e. desorp-
tion of physically adsorbed water occurs; or in the case of excess
acid on the mineral surface, the elimination of physically adsorbed or-
ganic component which can be removed by washing. In the second
temperature range of 200 °C to 400 °C, organic components that are Fig. 4. Comparative DTA curves of starting and calcite samples modified by “dry”
chemically adsorbed on the calcite surface are oxidized, as well as method.
212 S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216

preparation for modification using the “dry method”, the mineral was the added amount and the molecular weight of the stearic acid and
heated to 70 °C, so that part of the moisture was removed. This is sup- one calculated from the TG curve). Unlike the “wet modification”,
ported by smaller weight loss of CD-0.5 sample compared to the using this method, we suppose that the stearic acid is not adsorbed
starting sample of calcite, which was not heated. Thus, in this tempera- evenly i.e. that the second and multilayers form even before the sur-
ture range, a desorption of water samples with modified samples using face of the calcite is monolayer. This deviation is due to the unequal
the “dry method” is not a dominant reaction, however the weight loss distribution of stearic acid and domination of the physical sorption,
can be primarily attributed to the weakly bound acid or free stearic using the “dry method”.
acid. A part of DTA curves in Fig. 4 up to 200 °C is similar to the part For calcite modified by the “dry method”, the calculated surface
of DTA curve that represents pure stearic acid in this temperature area per molecule of stearic acid decreases with increasing concentra-
range [14]. tions of stearic acid used in the modification, from 0.540 nm 2 for
In Fig. 4, a wide exothermic peak of lower intensity (at around sample CD-0.5 to 0.063 nm 2 for sample CD-4.
120 °C) is observed at DTA curves of modified samples in the first Statistical monolayer (0.280 nm 2/molecule or 5.931 μmol/m 2)
temperature range (20–200 °C). This can be attributed to the weakly can theoretically be achieved at the concentration of stearic acid
bound (and/or free) molecules of stearic acid, which do not have a 0.81 phr ≈ 0.81%. However, according to the weight loss (Fig. 3)
mutual interaction between hydrocarbon chains due to the low con- and calculated degree of coverage, θ, (Table 3) in this experiment
centration. Constant weight loss at concentrations higher than it was achieved at a concentration of 1%. This signifies that statisti-
1.5 phr ≈ 1.5% of the used acid, indicates that the amount of such cally at higher concentrations of 1 phr ≈ 1% acid, in addition to the
spatially organized molecules in the samples of CD-2, CD-3 and monolayer, the formation of the second layer will also occur, i.e.
CD-4 is not increased. formation of multilayers.
In the calcite samples modified with fatty acids, the range of 200 to The calculated value of calcite surface area per molecule of stearic
400 °C is the area of organic component combustion — chemically acid of 0.54 nm 2 at the lowest concentration (0.5 phr of ≈ 0.5%)
adsorbed or bound stronger physically to the surface of minerals is pres- (Table 3) indicates that the alkyl chains of inhomogeneous adsorbed
ent. DTA diagrams of samples in this temperature range show that the organic compound are most likely arranged horizontally to the sur-
exothermic maximum occurs at temperatures above 310 °C. Increasing face of calcite. Statistical monolayer is formed at 1 phr acid, or the
the concentration of stearic acid during the modification process results surface area of 0.280 nm 2 per molecule of acid, indicating that the
in increasing of the intensity of this exothermic maximum, which is si- steric effect predominates in the “dry method”. In the “dry modifica-
multaneously moved to higher temperatures (from 311 °C for sample tion” the hydrocarbon chains at acid concentrations higher than
CD-0.5 to 342 °C for sample CD-4). In addition to this exothermic maxi- 0.5 phr, cannot be oriented perpendicularly to the mineral surface.
mum on DTA curves of modified samples, except for CD-0.5, another exo- Thus, they remain spatially arranged at an angle relative to the surface,
thermic maximum is observed at temperatures ≤ 251 °C for all modified i.e. with the hydrocarbon chains, as well as in oblique (gauche) confor-
calcite samples. The intensity of this maximum also increases with in- mation, with contribution of modification temperature. It can also lead
creasing of the concentration of stearic acid in the modification of calcite. to the spatial blocking of some primary surface centers, available for
Since the starting calcite does not show the weight loss in this chemical adsorption of the organic components. Therefore, the surfac-
temperature range, it is concluded that the weight loss derives from tant ions can generally be less chemically adsorbed in the “dry method”
the presence of the organic components strongly bound to the calcite than in the “wet modification method”. With further increase in the
surface and increases with increasing concentration of stearic acid, concentration of stearic acid in the modification method (c > 1 phr),
used in the modification method (from 0.42% for sample CD-0.5 to the trend of decline in the calculated value of calcite surface per mole-
3.60% for sample CD-4). cule of stearic acid continues (Table 3). This indicates that larger
The weight loss in the second temperature range (200 °C to 400 °C) amount of acid must be physically adsorbed on the calcite surface. Fur-
could also be used to calculate the statistical coverage of minerals [9,13], ther conformation of molecules (ions) in a condensed state also deter-
and to show the arrangement of stearic acid molecules on calcite sur- mines the interaction with the adjacent molecules (ions).
face. The mineral surface area per molecule of stearic acid was calculat- In regards to the “dry method” of sample preparation, total mass of
ed according to Eq. (4). Calculation of adsorbed organic compound was added acid remains in the sample, so it can be assumed that the disper-
made with an approximation that those are the only molecules of sion mechanism of the molten acid and its binding to the surface is some-
stearic acid. Rezaei et al. [13] reported that in the case of the “dry mod- what different from the dispersion mechanism in the “wet modification”.
ification method”, one molecule of stearic acid in the statistical mono-
layer occupies the calcite surface area of 0.280 nm2. The results are
presented in Table 3.
100
From Table 3 we can see the evident difference between the two
values of the adsorbed amount (expected one, calculated based on
80
Table 3
Active ratio, %

Adsorbed amount per mass unit and calcite surface area, surface area per molecule of
stearic acid and degree of coverage θ in samples modified by the “dry method”. 60

Sample Adsorbed Adsorbed Adsorbed Surface Degree of


amounta, amountb amount, area per coverage, θ 40
μmol/g , μmol/g μmol/m2 molecule, nm2

CD-0.5 17.6 14.76 3.075 0.540 0.52


CD-1 35.1 28.474 5.931 0.280 1.00 20
CD-1.5 52.7 43.59 9.081 0.183 1.53 dry
CD-2 70.2 61.52 12.817 0.130 2.16 wet
CD-3 105.3 92.45 19.260 0.086 3.25 0
CD-4 140.4 126.55 26.365 0.063 4.44 0 1 2 3 4
a
Adsorbed amount calculated mathematically based on the added amount and the Amount of stearic acid, %
molecular weight Γ = [m/M]/100, where m is the amount of the added stearic acid
(0.5, 1, 2, 3 and 4 g) and M is the molecular weight of the stearic acid (284.5 g/mol). Fig. 5. The effect of amount of stearic acid on the active ratio of obtained products, for
b
Adsorbed amount calculated from the weight loss. “dry and wet modification methods”.
S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216 213

For the calcite samples modified with stearic acid by the “dry 3.5. Mechanism of interaction between calcite and stearic acid in the “dry
method”, it can be generally said that the inhomogeneous adsorbed modification method”
local double layers are formed at lower concentrations of acid. How-
ever, at higher concentrations, there is a regrouping of such inhomo- In the “dry modification method” of calcite there is no formation of
geneous adsorbed local double layers and further physical adsorption, new active centers for chemisorption, as it happens during the “wet”
which leads to the formation of surface crystallized stearic acid. Due modification, and only the primary surface −Ca+ centers are on the
to this inhomogeneous adsorption, the major difference in the value calcite surface. Therefore, this paper assumes the mechanism of chemi-
of the coverage degree can be marked if we compare the results for sorption which involves the surface dissociation of molten acid mole-
both, “dry and wet methods”, presented in Tables 2 and 3. During cules and formation of calcium stearate and surface bicarbonate:
the “wet modification” the alkyl chains are evenly distributed. How-
ever, during the “dry method”, we have the formation of the second −Ca+ −CaOO(CH2)16CH3
and multilayer even before the formation of the first one is complete.
+ CH3(CH2)16COOH
With the same amount of stearic acid of 4%, degree of coverage is 2.29
and 4.4, respectively for the “wet and dry methods”.
−CO3− −CO3H

Fig. 5 shows the surface dissociation of molten stearic acid on


3.3. Floating test
calcite and possible structure of adsorbed layer on calcite (S =
4.8 m 2/g) after the “dry modification” at stearic acid concentrations
To investigate the effect of the amount of stearic acid at calcite sur-
c > 0.5 phr boundary between calcite and molten stearic acid. The
face on the hydrophobicity of obtained materials, the active ratio of
mechanism of surface dissociation of stearic acid is presumed where
each product was determined. These results are presented in Fig. 5.
H + ion moves to the surface carbonate ion (−) and stearic ion is
As can be seen from Fig. 5 the active ratio of 100% was achieved
chemisorbed on the primary surface center of Ca + ion. Only the pri-
with 1.5% and 3% of stearic acid using the “wet and dry modifica-
mary − Ca + centers are available for chemisorption.
tion”, respectively. This data is in agreement with the conclusions
Possible structure of the adsorbed layer presented in Fig. 8 shows
that we made according to the calculated surface area per mole-
the chemisorption of stearate due to the steric effect and oblique/
cule and degree of coverage. Since it is reported in the literature
gauche conformation of hydrocarbon chains. Acid molecules can be
that the higher the active ratio, the better modification effect is
physically adsorbed or bonded to the surface by hydrogen bonds, or
[6,14,20], the obtained results indicate that with the “wet modifi-
as dimers due to the interaction of hydrocarbon chains. Due to the
cation” complete hydrophobicity is achieved with smaller amount
steric effect, a part of surface Ca centers can be blocked, which is
of the stearic acid.
explained by the fact that the physical adsorption accounts for over
1.5%. The present molecules in the adsorbed layer are bound by differ-
ent strengths of connections, inhomogeneous, that was confirmed by
3.4. Microphotographs
the results obtained by thermal analysis.
Microphotographs of the starting calcite and calcites modified by
3.6. Mechanism of interaction between calcite and stearic acid in the
wet and dry modification methods with 1.5 and 3% of stearic acid
“wet modification method”
are presented in Figs. 6 and 7.
Microphotographs of starting sample (Figs. 6a and 7a) show inter-
Based on all test results [14], it is possible to assume the interac-
ference colors of very high order, which are characteristic for the
tion mechanism of calcite, water and stearic acid in the “wet modifi-
uncoated calcite. Compared to the starting sample in microphoto-
cation method” in two cases:
graphs of both wet-modified calcites (Fig. 6b and c) it is obvious
that in the presence of water as an immersion liquid, particles of coat- i) at concentration of stearic acid c b 1.5% (Fig. 9), and
ed calcites tend to concentrate, and thus agglomerates are clearly vis- ii) at concentration of stearic acid c > 1.5% (Fig. 10).
ible. These results may be another evidence that even with 1.5% of
stearic acid high hydrophobicity of coated product is achieved. Calcite is a salt type mineral, soluble in water, depending on
Microphotographs of dry modified calcite are presented in Fig. 7. dissolved CO2 and pH 13. Based on alkaline pH value of suspension
The microphotograph of the sample modified with the 1.5% of stearic (pH = 10.14), it can be concluded that the suspension is dominated
aid (Fig. 7a) clearly indicates the presence of the interference colors of by the reactions that lead to an increase in the concentration of hy-
very high order that imply the presence of the uncoated calcite. With droxyl ions, for example:
3% of stearic acid, calcite is coated and the agglomerates generate and
2þ − −
interference colors vanished and this is visible in Fig. 7c. CaCO3 þ H2 O⇄Ca þ HCO3 þ OH :

a b c

Fig. 6. Microphotographs of (a) starting calcite, (b) CS-1.5 and (c) CS-3; wet method.
214 S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216

a b c

Fig. 7. Microphotographs of (a) starting calcite, (b) CD-1.5 and (c) CD-3; dry method.

In this reaction the carbonate ion Ca + is formed as the primary Fig. 9 shows a possible structure of the Stern layer, where the chem-
center for adsorption. isorption occurs. Acid molecules can be bound by a hydrogen bond,
Fig. 9 shows the boundary between calcite and aqueous media. which occurs in the diffuse layer by physical adsorption process.
Due to the solubility of calcite in water, Ca + ions exist in the solution Fig. 10 shows a possible arrangement of chemical species at higher
and a calcium hydroxide CaOH is formed in very basic media that is concentrations when the physical adsorption dominates. Acid mole-
chemisorbed on the primary surface center of the carbonate ion Ca +. cules can be physically adsorbed as molecules and acid dimers that
Adsorption of OH ions may occur on the primary center Ca+ due to can be formed by decomposition of micelle under the influence of
very basic media. Subsequently, secondary surface centers are formed surface. Due to the solubility of calcite and increased concentrations
that can participate in the interaction of ions with stearic ions, formed of Ca + ion in the electric double layer, the surface precipitation of
by stearic acid dissociation. According to literature, the critical concentra- Ca-stearate can occur. Physically adsorbed species or molecules are
tion of micelle formation (CCM) of stearic acid in distilled water is from weaker or stronger bound in the adsorbed layer, which is also
4.5 × 10−4 mol dm−3 (at 20 °C) to 7 × 10−4 mol dm−3 (at 60 °C) influenced by the interaction between the hydrocarbon chains and
[20]. Adding the stearic acid at concentration above CCM, as is the case interactions in relation to the surface.
in this paper, in a strongly basic aqueous solution (pH = 10.14), leads In the “wet process” with increasing adsorption density, the adsorbed
to formation of micelles on one side, and dissociation of stearic acid of ions and molecules – due to their possible mutual interactions – especially
free molecules or molecules from micelle itself on the other. The between the hydrocarbon chains and due to the thickness of electrical
resulting stearic ions can be chemisorbed on the primary centers with
Ca+ ions or participate in the ion exchange reaction with OH− ions
from the secondary surface centers.
On the border of solid/liquid or mineral/solution, an electric dou-
ble layer is always formed consisting of the Stern and diffused layer.

Fig. 8. Surface dissociation of molten stearic acid on calcite and possible structure of
adsorbed layer on calcite (S = 4.8 m2/g) after “dry modification” at stearic acid con- Fig. 9. The assumed mechanism of interaction in the system calcite/water, stearic acid and
centrations c > 0.5 phr. structure of the Stern layer at stearic acid concentration c b 1.5 phr; (8.8 b pH b 10.14).
S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216 215

Fig. 10. Structure of the adsorbed layer (the Stern layer and a part of diffuse layer) in the system calcite/water, stearic acid at stearic acid concentration c > 1.5 phr (pH = 8.7–8.8).

double layer, achieve a vertical orientation relative to the surface and keep 2. In the “wet modification” a double electric layer was formed on the
the trans-conformation of hydrocarbon chains. boundary calcite/water, whose structure, i.e. hydrophilic properties,
were dependent on the concentration of potential determination
4. Conclusions ions, Ca2+ and CO32−, pH and adsorption of water molecules and
ions (and molecules) formed in a suspension of calcite due to its sol-
Based on the obtained results and their discussion, the following ubility in alkaline medium. The structure of such double electric layer
conclusions can be made: further influenced the interaction of minerals with anionic surfactant
(stearic ion and molecule of stearic acid). In the presence of stearic
1. In the “dry modification method”, only the primary surface centers of
acid, pH of the suspension decreased due to its dissociation in alkaline
−Ca+ are on calcite surface and there is no formation of the new ac-
medium and neutralization reaction. As the result of these reactions
tive centers for chemisorption, as during the “wet modification”.
during suspension, besides the undissociated acid molecules there
Based on this result, the chemisorption is the assumed mechanism
were also stearine ions CH3(CH2)16COO− and (St−) in a certain con-
where the surface molten acid molecules dissociate, and surface calci-
centration. Binding mechanism of stearine ions, and the remaining
um stearate and surface bicarbonate are formed.
undissociated molecules to the surface of calcite as a function of con-
Due to the steric effect of hydrocarbon chains of chemisorbed stearate
centration of used stearic acid had an impact on the degree of mineral
(their oblique conformation), a part of the primary centers is blocked,
hydrophobicity, or the removal of chemisorbed H+ and OH− ions
which affects the degree of surface dissociation, or the amount of the
from the Stern part of the double electrical layer.
chemisorbed organic component. Statistical monolayer in the “dry
modification” (0.280 nm2/molecule or 5.931 μmol/m2) is theoretically At concentration of stearic acid up to 1.5 phr, the chemically adsorbed
achieved at stearic acid concentration of 0.81%. Calculated coverage de- organic component (stearic ions) dominates. This can be explained by for-
gree θ = 1.00 at acid concentration 1 phr ≈ 1% indicates that at any mation of double electric layer on the boundary calcite/water or heteroge-
higher concentrations – statistically – the formation of the bilayer neous surface of calcite, which in addition to the primary surface −Ca+,
and multilayer starts. However, two exothermic peaks are observed has the new active centers, for example \CaOH, formed by protonization
on DTA curves, indicating that the organic phase is adsorbed both or \CO3CaOH, formed by adsorption of CaOH+ on \CO3−. Therefore it is
chemically and physically on this sample. This generally indicates the possible that, in addition to the stearate chemisorption on −Ca+ centers,
inhomogeneity of modified calcite surface, i.e. physical adsorption the reaction of ion exchange also occurs between St− from solution and
through partly formed chemisorbed layer (formation of inhomoge- OH− from the newly-formed surface centers, which increases the amount
neous adsorbed local double layers). of chemisorbed SAM.
At higher concentrations of stearic acid in a liquid state — as monomers At higher adsorption density of stearate, the interaction between
and dimers, they have to be “arranged” to form the multiple layers in the hydrocarbon chains occurs, which results in their vertical orienta-
which the conformation of hydrocarbon chains is still oblique, but it tion in relation to the surface of minerals and larger surface arrange-
is determined by the intensity of mutual interactions. Cooling on the ment. Statistical monolayer (0.165 nm 2/molecule or 10.064 μmol/m 2
surface of minerals through chemisorbed stearate, leads to a “new” of calcite sample with specific surface area 4.8 m 2/g) can be achieved
solid phase to be formed from such “organized” stearic acid. Full hy- theoretically at concentration of 1.37 phr ≈ 1.36% stearic acid. In the
drophobic surface at higher concentrations, 3 and 4 phr, indicates modified calcite at concentration 1.5 phr of acid, the calculated statis-
that the last layer of molecules of surface condensed stearic acid orien- tical coverage of mineral θ = 1.05 has indicated that statistically only
tates with hydrocarbon chains opposite of mineral surface (i.e. the acid at this acid level the second layer starts to be modified in addition to
on calcite is organized as a surface lamellar micelle). the monolayer.
216 S.R. Mihajlović et al. / Powder Technology 245 (2013) 208–216

At higher concentrations of 1.5 phr stearic acid, in addition to the [7] V. Kovačević, S. Lučić, D. Hace, A. Glasnović, I. Šmit, M. Bravar, Investigation of the
influence of fillers on the properties of poly(vinyl acetate) adhesives, Journal of
chemisorbed stearate, a large amount of physically adsorbed organic Adhesion 47 (1994) 201–215.
component is detected. It consists of stronger or weaker bound stearic [8] A. Martínez-Luèvanos, A. Uribe-Salas, A. Lόpez-Valdivieso, Mechanism of adsorp-
acid molecules to the surface as well as the molecules of surface pre- tion of sodium dodecylsulfonate on celestite and calcite, Minerals Engineering 12
(1999) 919–936.
cipitated calcium stearate, CaSt2, formed in a double electric layer of [9] G.K.A. Rezaei, A.A. Hamouda, Effect of fatty acids, water composition and pH on
calcite. The used concentrations of stearic acid were higher than the the wettability alteration of calcite surface, Journal of Petroleum Science and En-
critical concentration of micelle formation (CCM), so that the physi- gineering 50 (2006) 140–150.
[10] D. Holclajtner-Antunović, General Course of Physical Chemistry, Zavod za udžbenike
cally adsorbed molecules were free acid molecules in a solution i nastavna sredstva, Belgrade, Serbia, 2000.
and/or molecules from the formed micelles in a solution and then [11] D. Vučinić, S. Popov, Physical Chemistry, Faculty of Mining and Geology, Belgrade,
destroyed under the influence of mineral surface. Serbia, 2004.
[12] V. Dondur, Chemical Kinetics, Faculty of Physical Chemistry, Belgrade, Serbia,
1992.
Acknowledgments [13] G.K.A. Rezaei, R. Denoyel, A.A. Hamouda, Wettability of calcite and mica modified
by different long-chain fatty acids (C18 acids), Journal of Colloid and Interface
This paper is the result of research on the Project TR 34013 titled “De- Science 297 (2006) 470–479.
[14] S. Mihajlović, Ž. Sekulić, A. Daković, D. Vučinić, V. Jovanović, J. Stojanović, Surface
velopment of technological processes for obtaining of ecological materials properties of natural calcite filler treated with stearic acid, Ceramics-Silikaty 53
based on nonmetallic minerals”, and the Project TR 34006 titled “Mecha- (2009) 268–275.
nochemical treatment under insufficient mineral resources”, funded by [15] Y. Sheng, B. Zhou, J. Zhao, N. Tao, K. Yu, Y. Tian, Z. Wang, Influence of octadecyl
dihydrogen phosphate on the formation of active super-fine calcium carbonate,
the Ministry of Education, Science and Technological Development of Journal of Colloid and Interface Science 272 (2004) 326–329.
the Republic of Serbia for the period 2011–2014. [16] E.J. Sullivan, D.B. Hunter, R.S. Bowman, Topological and thermal properties of
surfactant-modified clinoptilolite studied by tapping-mode atomic force micros-
copy and high-resolution thermogravimetric analysis, Clays and Clay Minerals 1
References (1997) 42–53.
[17] W. Fuerstenau, P. Pradip, Zeta potentials in the flotation of oxide and silicate min-
[1] W. Wu, S.C. Lu, Mechano-chemical surface modification of calcium carbonate par-
erals, Journal of Colloid and Interface Science 114–115 (2005) 9–26.
ticles by polymer grafting, Powder Technology 137 (2003) 41–48.
[18] E.H.M. Wright, N.C. Pratt, Solid/solution interface equilibria for aromatic mole-
[2] P. Rungruang, B.P. Grady, P. Supaphol, Surface-modified calcium carbonate parti-
cules adsorbed from non-aromatic media. Part 2. Aromatic carboxylic acids, Jour-
cles by admicellar polymerization to be used as filler for isotactic polypropylene,
nal of the Chemical Society, Faraday Transactions 74 (1974) 1461–1471.
Colloids and Surfaces A 275 (2006) 114–125.
[19] G. Hansen, A.A. Hamouda, R. Denoyel, The effect of pressure on contact angles and
[3] D.S. Keler, P. Luner, Surface energetics of calcium carbonates using inverse gas
wettability in the mica/water/n-decane system and the calcite + stearic acid/
chromatography, Colloids and Surfaces A 161 (2000) 401–415.
water/n-decane system, Colloids and Surfaces A 172 (2000) 7–16.
[4] E. Papirer, J. Schultz, C. Turchi, Surface properties of a calcium carbonate filler
[20] Y.F. Yang, X.F. Wu, G.Sh. Hu, B.B. Wang, Effects of stearic acid on synthesis of mag-
treated with stearic acid, European Polymer Journal 20 (1984) 1155–1158.
nesium hydroxide via direct precipitation, Journal of Crystal Growth 310 (2008)
[5] M.A. Osman, U.W. Suter, Surface treatment of calcite with fatty acids: a structure
3557–3560.
and properties of the organic monolayer, Chemistry of Materials 14 (2002)
4408–4415.
[6] Y. Sheng, B. Zhou, Ch. Wang, X. Zhao, Y. Deng, Z. Wang, In situ preparation of hy-
drophobic CaCO3 in the presence of sodium oleate, Applied Surface Science 253
(2006) 1983–1987.

Anda mungkin juga menyukai