You are on page 1of 14

379

Adsorption of Atenolol on Talc: An Indication of Drug Interference with


an Excipient
Zhaohui Li1,2,3,*, Nicole M. Fitzgerald2, Guocheng Lv1,*, Wei-Teh Jiang3 and Limei Wu1
(1) Beijing Key Laboratory of Materials Utilization of Nonmetallic Minerals and Solid Wastes, National Laboratory of
Mineral Materials, School of Materials Science and Technology, China University of Geosciences, Beijing, 29 Xueyuan
Road, Beijing, China 100083. (2) Geosciences Department, University of Wisconsin–Parkside, Kenosha, WI 53144, USA.
(3) Department of Earth Science, National Cheng Kung University, 1 University Road, Tainan, Taiwan 70101.

(Received date: 20 Jan 2015; Accepted date: 23 Mar 2015)

ABSTRACT: Talc is commonly used as an excipient for drug formulations. The


general expectation is that the excipient should have only minimal interactions
with the carrying drug. In this study, the adsorption of atenolol (AT), a β-blocker,
on talc, a clay mineral of pH-dependent surface charge, was evaluated under
different physicochemical conditions such as the initial AT concentration,
equilibrium time, solution pH, ionic strength and temperature. Our experiments
showed that talc had an AT adsorption capacity of 11 mmol/kg. In addition,
adsorption of AT on talc was instantaneous, which suggests that the adsorption
sites were located on the external surfaces or edges. Factors such as solution pH,
ionic strength and temperature all had minimal influence on AT adsorption,
although the adsorption process was exothermic and the free energy of adsorption
was negative, indicating weak physical adsorption. The Fourier transform
infrared results showed blue shifts of the bands corresponding to δ(C–OH) at
1410 and 1043 cm–1, indicating hydrogen bonding for the uptake of AT on talc
surfaces or edges.

1. INTRODUCTION

Beta-blockers (β-adrenergic receptor antagonists) are a class of widely prescribed cardiovascular


drugs, used for the treatment of hypertension, cardiac arrhythmias, cardioprotection after heart
attacks and anxiety disorders (Barbieri et al. 2012). Experimental studies of 11 β-blockers
indicated very low gastric absorption rate for these drugs; however, their intestinal and colonic
absorption rates were of the same magnitude between 0.12 and 4.28 hour–1 (Vilà et al. 1992).
Hydrophobicity is an important predictor of β-blocker adsorption; propranolol is the most
hydrophobic β-blocker studied, which was adsorbed to the greatest extent (Kibbey et al. 2007).
Atenolol (AT) is one of the most commonly prescribed β-blockers with a low hydrophobicity or
lipophilicity with a log KOW of only 0.1, suggesting a low affinity for solid substrates (Küster
et al. 2010).
Because of their widespread use and limited human metabolism, β-blockers are widely detected
in sewage effluents and surface waters (Küster et al. 2010). As such, a few recent studies were
conducted to investigate the interactions between selected β-blockers and clay minerals or
sediments (Küster et al. 2010; Rakić et al. 2013; Schaffer et al. 2012; Seema and Datta 2013;
Yamamoto et al. 2009). In addition, for the purpose of drug removal and evaluating antidotal

*Authors to whom all correspondence should be addressed. E-mail: li@uwp.edu (Z. Li); guochenlv@cugb.edu.cn (G. Lv).
380 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

effect, adsorption of AT on activated charcoal (AC) or granular activated carbon (GAC) was also
investigated (Rao et al. 2013; Sancho et al. 2012).
Talc is a common excipient in drug formulation (Hansen et al. 2011). In addition, talc has also
been used as a glidant in concentrations of approximately 1–2% by weight to improve the
flowability of the powder and as an antiadherent (Alderborn 2013). The general expectation is that
the excipient should have only minimal interactions with the carrying drug. As an excipient, talc
did not show any interaction with celecoxib, which is a non-steroidal anti-inflammatory drug
(Bozdağ-Pehlivan et al. 2011). However, strong ciprofloxacin (CIP) adsorption on talc was
reported (adsorption capacity, 0.74 mg/g; Ibezim et al. 1999). A modest binding of CIP to talc was
observed, which was attributed to the weak van der Waals forces and ion-exchange mechanism
(Poudel and Thapa 2013). A study by Ashford (2013) showed that talc interfered with the
absorption of cyanocobalamin owing to its ability to adsorb this vitamin. Adsorption of
zearalenone from synthetic gastric fluid and synthetic body fluid seemed to be less likely because
of the interaction between partial negative charges of zearalenone surface and edge hydroxyl
groups of the talc (Sprynskyy et al. 2012).
On a different perspective, the adsorption of drugs with poor water solubility on talc could
improve the drug dissolution rate and desorption, allowing for complete release of drugs (Jadhav
et al. 2013). Agglomerates prepared by combining ibuprofen and talc could extend ibuprofen’s
release time by up to 13 hours (Pawar et al. 2004). Talc facilitated the release of salicylic acid at
a faster rate in vitro, compared with stearate (Jadhav et al. 2013). Talc had been used in many
commercial antacids for stomach and indigestion problems. Talc has high resistance to acids and
chemicals, which makes it a good filler for tablets (Jadhav et al. 2013).
Because talc is an important excipient and AT is a typical β-blocker, we chose these compounds
to evaluate their interactions in aqueous suspension in terms of AT uptake as a function of initial
AT concentrations, solution pH, ionic strength, and contact time. The results of this study could
be used to guide excipient selections to minimize interactions.

2. MATERIALS AND METHODS

Talc was purchased from Acros (US). It has a mean particle size less than 200 mesh (<0.075 μm)
and an external surface area of 2.3 m2/g, measured by a multipoint BET method. Its chemical
formula is Mg3Si4O10(OH)2. It has approximately 2% of clinochlore, as determined by
semiquantitative X-ray diffraction (XRD) analyses. Talc has no permanent surface charges due to
its low isomorphous substitution in the tetrahedral and octahedral sites (Bergaya and Lagaly
2006), resulting in an extremely low cation-exchange capacity (CEC). The reported point of zero
charge was 7.7 (Burdukova et al. 2007; Lv et al. 2014).
AT (4-[2′-hydroxy-3′-[(1-methylethyl)amino]propoxy]-benzeneacetamide; Figure 1) was
purchased from Tokyo Chemical Industry Co, Ltd (Japan). It has a purity over 98%, a formula
mass of 266.336 g/mol, a pKa value of 9.6 (Küster et al. 2010; Manallack 2007; Silveira et al.
2014) and a water solubility of 9.34 g/l (Vinnakota et al. 2011). Experimental and predicted
solubilities of log S0 = –1.3 and –1.94, corresponding to 13 and 3 g/l, were also reported (Avdeef
et al. 2000). At 37 °C its solubility was 37 g/l, which is significantly higher than that at room
temperature (Hapeshi et al. 2010). The AT solution prepared using Milli-Q water was stable under
room temperature (Yamamoto et al. 2009). In all experiments, 1 g talc and 10 ml AT solution were
added to each 50-ml centrifuge tube and shaken at 150 rpm for 24 hours, except for the kinetic
study. The mixture was centrifuged at 3500 rpm for 10 minutes, and the supernatants were passed
Adsorption of Atenolol on Talc 381

O3

O1 C5 N2
C4 C14
O2
C10
C6
C1 C9 C11 C12
C3
N1
C2 C7
C8 C13

Chiral center

Figure 1. Molecular structure of atenolol. It could be a racemic mixture and the chiral centre is at the C marked with an
arrow. The red is O, the blue is N, the dark grey is C and the white is H. The numbers followed those of Castro et al.
(2007a).

through 0.45-μm syringe filters before being analyzed for the equilibrium AT concentrations using
a UV–Vis method.
The initial AT concentrations were 0.0, 0.2, 0.6, 0.8, 1.0, 1.2, 1.6 and 2.0 mmol/l for the AT
adsorption isotherm study, and was fixed at 2.0 mmol/l for all other studies. The equilibrium time
was 0.25, 0.5, 1, 2, 4, 8, 16 and 24 hours for the AT adsorption kinetic study. For the pH-dependent
study, the equilibrium solution pH was maintained at 2–11 with an increment of 1 with periodic
adjustments. The ionic strength was adjusted to reach NaCl concentrations of 0.001, 0.01, 0.1 and
1.0 M when preparing solutions. The temperature study was conducted at 22, 32, 42 and 52 °C.
A UV–Vis method at a wavelength of 274 nm was used to analyze equilibrium AT concentrations
(Sable et al. 2012; Weich et al. 2007). At this wavelength, the absorbance value of AT was stable
at pH 1–11 (Seema and Datta 2013). The Fourier transform infrared (FTIR) analysis was conducted
using a JASCO FT/IR-4100 spectrometer equipped with a ZnSe crystal and attenuated total
reflection accessory. The spectra were obtained from 600 to 4000 cm–1 by accumulating 256 scans
at a resolution 4 cm–1. Powder XRD analyses of the raw and AT-adsorbed talc were carried out with
Cu-Kα radiation at 40 kV and 40 mA on a D8 ADVANCE diffractometer (Bruker). Unoriented
samples were scanned from 2θ = 2 to 12° with a scanning speed of 0.01°/second.
Molecular simulation was performed under the ‘Forcite’ module in the Materials Studio 6.0
software to investigate the adsorption sites of AT on the (010) surface of talc. The talc model was
constructed and the atomic coordinates were derived from the space group of C2/C with a = 5.26 Å,
b = 9.10 Å, c = 18.81 Å, α = γ = 90° and β = 100°. The supercell of the model was made of 12 talc
unit cells at 3a × 2b × 2c. On the surface of XYZ, the volume was 15.78 Å × 18.20 Å × 37.62 Å.
The established model was optimized geometrically. The temperature was set at 298 K and time was
1 ns, with a time step of 1 fs. The ClayFF force field was used during simulation. The Ewald
summation method was used to calculate the interaction. After the system reached equilibrium, the
NVT kinetic simulation was performed under the same time constant and temperature conditions.
The data were collected on the last 300 ps for further analyses.

3. RESULTS AND DISCUSSION


3.1. AT Adsorption Isotherm

Several adsorption isotherm models were used to fit the experimental data and the Langmuir
adsorption isotherm best fitted with the data [Figure 2(a)]
382 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

(a)
8

Amount of AT adsorbed (mmol/kg)


6

0
0.0 0.4 0.8 1.2 1.6
Equilibrium AT concentrations (mmol/l)

(b)
6
Amount of AT adsorbed (mmol/kg)

4 6.0
t/qt (h kg/mmol)

5.0
4.0
3.0
2.0 y = 0.20x + 0.3
2 1.0 r2 = 0.9996
0.0
0 5 10 15 20 25 30
Time (h)

0
0 5 10 15 20 25
Time (h)

Figure 2. (a) Adsorption isotherm of atenolol on talc. The line is the Langmuir fit to the observed data. (b) Kinetics of AT
adsorption on talc. The line represents pseudo-second-order fits to the observed data.

K L Sm C L (1)
CS =
1 + K L CL

where CS is the amount of AT adsorbed at equilibrium (mmol/kg), Sm is the apparent adsorption


capacity (mmol/kg), CL is the equilibrium AT concentration (mmol/l) and KL is the Langmuir
coefficient (l/mmol). The fitted AT adsorption capacity on talc was 11 mmol/kg, or 3.0 mg/g,
which is much higher than that of CIP on talc at 0.74 mg/g (Ibezim et al. 1999), and the fitted KL
value was 0.75 l/mmol. These values are, however, lower than the value (20 mmol/kg) achieved
for AT adsorption on kaolinite and bentonite (Rakić et al. 2013). The uptake of AT on natural
kaolin was attributed to multilayer adsorption, which was indicated by a slight increase in AT
adsorption values in the last part of the isotherm (Rakić et al. 2013). By contrast, a linear
adsorption was observed for AT uptake on some river sediments or on a silt loam soil, with
reported Kd values ranging from 1.3 l/kg for the river sediments to 110 l/kg for the silt loam soil
Adsorption of Atenolol on Talc 383

due to its higher CEC value (Yamamoto et al. 2009). The equilibrium uptake of AT on
montmorillonite also followed the Langmuir model with an AT adsorption capacity of 86 mg/g, or
320 mmol/kg (Seema and Datta 2013). The adsorption of AT on GAC resulted in a capacity of
750 mmol/kg, compared with the measured basicity of 802 mmolc/kg (Sancho et al. 2012).
AT has a surface area of 34 Å2 (Seelig et al. 1994). To achieve the adsorption capacity of
11 mmol/kg, the total external surface area needed was 2240 m2. This value is approximately the
same as the external surface area of talc (2.3 m2/g). Thus, the adsorption of AT is limited to the
external surfaces of talc with monolayer adsorption, confirming the use of the Langmuir equation
for fitting the experimental data.

3.2. AT Adsorption Kinetics

Kinetically, the uptake of AT on talc was very fast and an equilibrium could be achieved in just
2 hours [Figure 2(b)]. Similarly, AT adsorption reached equilibrium in less than 2 hours on AC
(Rao et al. 2013) or less than 8 hours on GAC (Sancho et al. 2012). The quick uptake of AT on
talc again indicated that the adsorption sites were on the external surface or edges. Several kinetic
models were used to fit the experimental data and the pseudo-second-order model fitted the
experimental data well

kq 2e t (2)
qt =
1 + kq e t

where qe and qt are amounts of AT adsorbed at equilibrium and at time t (mmol/kg), k is the
pseudo-second-order rate constant (kg/mmol hour). A rate constant of 1.3 kg/mmol hour, and
an initial rate of 33 mmol/kg hour were obtained for AT uptake on talc from an initial
concentration of 2 mmol/l. In comparison, a rate constant of 0.082 g/mg minute was reported for
AT adsorption on montmorillonite (Seema and Datta 2013). Rate constants were 0.1 and
1.9 kg/mmol hour and the initial rates were 68 and 476 mmol/kg hour, respectively, for AT
adsorption on kaolinite from initial concentrations of 4 and 2 mmol/l (Hu et al. 2015).

3.3. Effect of pH

Overall, the influence of equilibrium solution pH on AT adsorption on talc was not significant,
except at pH 11, where the AT adsorption was 20–30% lower [Figure 3(a)]. Similarly, relatively
stable adsorption of propranolol on attapulgites or modified attapulgites was achieved at the initial
pH range of 3–11 (Deng et al. 2011). By contrast, AT adsorption on AC was significantly affected
by solution pH with the highest AT adsorption occurring at neutral pHs (Rao et al. 2013). For
kaolinite, the maximum AT uptake was at pH 6–9 (Hu et al. 2015). An opposite trend was also
observed for AT adsorption on a sandy aquifer material, which decreased from 23.3 to 15.8 as the
pH increased from 4 to 8 (Schaffer et al. 2012).

3.4. Solution Ionic Strength Effect

The influence of solution ionic strength on AT adsorption on talc was also minimal. The amount
of AT adsorbed varied from 2.8 to 3.1 mmol/kg at background NaCl concentrations of
0.001–1 mol/l [Figure 3(b)]. The free energy of adsorption ΔG can be calculated using the
following formula:
384 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

(a)
7

Amount of AT adsorbed (mmol/kg)


6

0
2 4 6 8 10 12
Equilibrium solution pH

(b)
3.5 –3.0
Amount of AT adsorbed (mmol/kg)

3.3 –3.2

∆G (kj/mol)
3.1 –3.4

2.9 –3.6

2.7 –3.8

2.5 –4.0
–3 –2 –1 0
Log (NaCl)

Figure 3. AT adsorption on talc as affected by (a) equilibrium solution pH and (b) ionic strength.

ΔG = –RTlnKd (3)

where Kd = CS/CL, R is the gas constant and T is the temperature in K. The ΔG calculated was in
a narrow range of –3.3 to –3.7 kJ/mol for AT adsorption under different ionic strength conditions.
This result is different from AT adsorption on kaolinite where a significant reduction in AT
adsorption was found as the ionic strength increased from 0.001 to 1 mol/l (Hu et al. 2015).
Propranolol molecules could be easily desorbed into the aqueous phase as the Na+ concentration
of the solution increased (Deng et al. 2011).

3.5. Effect of Temperature

Temperature also had little effect on AT adsorption on talc, although a slight decrease in Kd was
found as the temperature increased, reflecting a slightly exothermic reaction. The change in
enthalpy (Table 1) after AT adsorption was only –7.5 kJ/mol, confirming the exothermic reaction,
Adsorption of Atenolol on Talc 385

TABLE 1. Thermodynamic Parameters of AT Adsorption on Talc from an Initial Concentration of 2 mmol/l


ΔG ΔH ΔS
(kJ/mol) (kJ/mol) (kJ/mol K)

T (K) 295 305 315 325


–3.1 –3.0 –2.8 –2.7 –7.5 –0.01

whereas the ΔG values were between –2.7 and –3.1 kJ/mol, which are agreeable with the results
obtained from the ionic strength test. The extremely small values of ΔG also supported weak
physical adsorption. The entropy change was –0.01 kJ/mol K. A similar exothermic adsorption of
AT on kaolinite had slightly more negative ΔG values in the range of –7 to –9 kJ/mol and more
negative ΔH values in the range of –25 to –14 kJ/mol (Hu et al. 2015). By contrast, AT adsorption
on AC was endothermic, with ΔG, ΔH and ΔS values of –2 to –5 kJ/mol, 29 kJ/mol and
0.09 kJ/mol/K, respectively (Rao et al. 2013).

3.6. Fourier Transform Infrared Analysis

The FTIR spectra of crystalline AT, raw talc and talc adsorbed with different amounts of AT
showed the characteristic bands of these substances [Figure 4(a)]. The band locations and
assignments for AT are presented in Table 2. The bands pertaining to talc are not discussed,
however. The band at 3364 cm–1 was attributed to the ν(–OH) vibration, whereas that at 3350 cm–1
was attributed to the νa(NH) vibration (Castro et al. 2007a). In this study, the νa(NH) vibration
shows a single peak at 3360 cm–1, which is embedded into the –OH vibration of talc after AT
adsorption. However, the peak intensity increased as the input AT concentration increases,
suggesting participation of the –OH group on AT adsorption. The band at 2972 cm–1 originated
from C–H stretching (Seema and Datta 2013) or alcoholic –OH group (Shirsand et al. 2011), but
after AT adsorption on talc, this band appears at 2975 cm–1 (Table 2).
In the 1800–1550 cm–1 range in the spectra of (R,S)-AT, there were four component bands. The
1585 and 1612 cm–1 bands correspond to the C=C ring stretch vibrations, whereas the 1671 and
1639 cm–1 bands were components of amide I (Castro et al. 2007a). The intensity of the band
at 1638 cm–1, originating from C=O valence vibrations (amide band I), decreased significantly
and shifted to 1668 cm–1 after inclusion into the cavity of 2-hydroxypropyl-β-cyclodextrin,
which restricted its vibration (Nikolic et al. 2007). It shifted to 1650 cm–1 after the interaction of
AT with pseudoboehmite (Novickis et al. 2013). In this study, this band disappeared completely
after AT adsorption on talc [Figure 4(b)]. The band at 1633 cm–1 was attributed to the C=C ring
(Castro et al. 2007a) in crystalline AT, and this band was also not observed after AT adsorption
on talc.
The band at 1240 cm–1 is attributed to the alkyl aryl ether linkage (Seema and Datta 2013;
Shirsand et al. 2011). For the crystalline AT in this study, this appears as a single sharp peak at
1237 cm–1. This band splits into two peaks, located at 1230 and 1252 cm–1, respectively, after
AT adsorption onto the surface of kaolinite (Hu et al. 2015). However, following AT adsorption
on talc, this sharp band also disappeared [Figure 4(b)]. The band at 1410 cm–1 for crystalline AT
was attributed to δ(C–OH) (Castro et al. 2007a) and it blue shifted to 1460 cm–1 after being
adsorbed on talc (Table 2), suggesting the formation of hydrogen bonds during AT adsorption.
Hydrogen bonding gave rise to a red shift of ν(C = O) and a blue shift of δ(NH2). Partial
386 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

(a)

0.0

0.2

0.6

0.8

1.0

1.2

1.6

2.0

100

AT
80
Transmittance (%)

60

40

20
4000 3500 3000 2500 2000 1500 1000 500
Wavenumber (cm−1)

(continued)

overlapping of these two bands would make the spectra appear as a doublet as observed in the
case of (R, S)-AT (Castro et al. 2007a). Thus, hydrogen bonding may play a significant role in
AT uptake on talc. The band at 1043 cm–1 originated from the asymmetric alkyl aryl ether
linkage (Seema and Datta 2013) or δ(C–OH) (Castro et al. 2007a), and it also blue shifted to
1056 cm–1 after AT adsorption on talc. This again suggests hydrogen bonding interactions
between AT and talc. The bands with wavenumbers below 1000 cm–1 were excluded from
further discussion in this study.
Adsorption of Atenolol on Talc 387

(b)

0.0

0.2

0.6

0.8

1.0

1.2

1.6

110
2.0

100

90
Transmittance (%)

80

AT
70

60

50

40
1700 1600 1500 1400 1300 1200 1100 1000
Wavenumber (cm−1)

Figure 4. FTIR spectra of crystalline AT and talc with different amounts of AT adsorption in the range of (a) 600–4000 cm–1
and (b) 1000–1700 cm–1. The numbers are initial AT concentrations in mmol/l.

3.7. X-Ray Diffraction Analyses

The XRD peaks of AT are located at 2θ of 3.2°, 6.4°, 9.6°, 12.9°, 16.1° and 19.3°. These peaks
may correspond to the (200), (400), (600), (800), (1000) and (1200) diffractions [Figure 5(a)], as
the racemic (R, S)-AT had a space group of C2/c with a = 55.83 Å, b = 5.559 Å, c = 9.734 Å and
β = 100.042° (Castro et al. 2007b). The raw talc contained approximately 2% clinochlore as
evidenced by its (001), (002), (003) and (004) diffractions peaks, which occurred at 2θ = 6.2°, 12.44°,
18.84° and 25.06°. However, only the (001) peak is enlarged in Figure 5(b). After adsorption of
388 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

TABLE 2. FTIR Bands for AT and AT-Adsorbed on Talc


0.0 0.2 0.6 0.8 1.0 1.2 1.6 2.0 Band
mmol/l mmol/l mmol/l mmol/l mmol/l mmol/l mmol/l mmol/l Pure AT assignments

3359 3359 3359 3359 3359 3359 3359 3359 3358 ν(–OH) vibrationa
— — — — — — — — 3346 νa(NH) vibrationa
2975 2975 2975 2975 2975 2975 2975 2972 Asymmetric C–H
stretchingb or
alcoholic
–OH groupc
2927 2927 2927 2927 2927 2927 2927 2927 2921
1633 ν(C=C) ringa
1613 δ(NH) vibrationa
1583 ν(C=C) ringa
1515 Benzene ring
skeletal vibrationb,d
1457 1457 1458 1458 1458 1460 1460 1410 δ(C–OH)a
1385 1385 1385 1385 1385 1385 1385 1385 1381
— — — — — — — — 1338 ν(C–N), δ(NH)
vibrationa
— — — — — — — — 1301
— — — — — — — — 1237 Asymmetric alkyl
aryl etherb,c;
δ(C–OH)a
— — — — — — — — 1178 ν(C–N)a
1093 1093 1093 1093 1093 1091 1091 1088
— — — — — 1086 1086 1075 δ(C=C) ring out
of phasea
1056 1056 1056 1054 1054 1052 1050 1038 Asymmetric alkyl
aryl ether linkageb;
δ(C–OH)a
a
Castro et al. (2007a).
b
Seema and Datta (2013).
c
Shirsand et al. (2011).
d
Nikolic et al. (2007).

different amounts of AT, no change in XRD patterns of talc was observed (Figure 5), suggesting that
the adsorption of AT occurred only its external surfaces or edges. Moreover, the AT peaks were non-
detectable after mixing or adsorption on talc. Similar results were observed in the XRD patterns of
physical mixture of AT and hydrophilic swellable polymers used as excipients and their formulations,
indicating that the AT was completely dispersed in the solid substrate (Kumar et al. 2013).

DISCUSSION

Although studies of AT adsorption on different adsorbents were previously conducted,


contradicting results were reported. Cation exchange was suggested to play a dominating role for
the retardation of AT activity in different sediments (Schaffer et al. 2012; Yamamoto et al. 2009).
However, kaolinite and bentonite had almost the same AT adsorption capacity despite their drastic
Adsorption of Atenolol on Talc 389

(a)
35,000
24000
(200) (400) Talc
AT peak
30,000 18000

12000
25,000 (600)
6000
Intensity (cps)

(800)
20,000 0
2 4 6 8 10 12 14
2θ (degrees) mix
15,000
2.0
10,000 1.0

5000 Clinochlore peak 0.0

0
2 4 6 8 10 12
2θ (degrees)

(b)

Clinochlore
500
mix
400

300 2.0

200 1.0

100
0.0

0
3 4 5 6 7
2θ (degrees)
Figure 5. (a) X-ray diffraction patterns of crystalline AT (inset), raw talc and talc adsorbed, or mixed, with different
amounts of AT and (b) their enlargement at 3–7° to show the trace amount of clinochlore.

difference in the CEC values (Rakić et al. 2013). A weak AT adsorption on sediments suggested
that there is only a small contribution of electrostatic interactions to the overall AT adsorption,
whereas the organic carbon in the sediments was thought to have played an important role in AT
uptake (Küster et al. 2010). In a different study, the organic matter in the sediments was
considered playing a much smaller role in AT adsorption (Niedbala et al. 2013).
In this study, the AT adsorption capacity was 11 mmol/kg, and the CEC of the talc was less than
10 mmolc/kg (Spark 2003). Thus, initially it appeared as though the AT adsorption capacity closely
matched with the CEC value. However, results of ion chromatography analyses did not show any
detectable desorption of exchangeable cations accompanying AT adsorption from different initial
concentrations (data not shown), excluding cation exchange as a major mechanism for AT uptake
390 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

on talc. Moreover, the equilibrium solution pH was between 9 and 10, whereas the pKa value of
AT is 9.6. Thus, under equilibrium adsorption conditions, significant amount of adsorbed AT is in
the neutral molecular form. Furthermore, the solution pH only had minimal influence on AT
uptake in the pH range of 3–10 [Figure 2(a)], which further suggests minimal contribution of
electrostatic interactions. Lastly, the fractional organic carbon content of talc is essentially zero,
ruling out the contribution of organic matter to AT adsorption.
The surface oxygen atoms of talc are non-polar due to their participation in silicon–oxygen
tetrahedral structure; thus, the talc basal surface oxygen atoms are not ideal hydrogen bonding
sites (Du et al. 2012). However, due to the negative charges that these oxygen atoms carry, they
can still form weak hydrogen bonding when compared with typical hydrogen bonding (Du et al.
2012). Overall, an AT molecule has four hydrogen acceptors and three hydrogen donors with a
topographic polar surface area of 85 Å2 (Ertl et al. 2000). The blue shifts of the bands
corresponding to δ(C–OH) at 1410 cm–1(Castro et al. 2007a) and the alkyl aryl ether linkage
(Seema and Datta 2013) or δ(C–OH) (Castro et al. 2007a) at 1043 cm–1 may suggest hydrogen
bonding for the uptake of AT on talc surfaces or edges. The –OH attached to the chiral centre
(O3 position) and the –O– linkage (O2 position) could form hydrogen bonding with O on the
tetrahedral sheet and/or on the broken bond of talc edges. The disappearance of C=O band
vibration at 1633 cm–1 may suggest the breakup of the C=O bond (O1 position) as well as the
formation of hydrogen bond with talc. Hydrogen bonding was an important factor for the
adsorption of carboxymethyl cellulose, a polysaccharide, on talc (Wang and Somasundaran 2005).
Molecular simulation showed close distance between the (010) surface of talc and the ring of AT,
indicating possible π-interactions or hydrophobic interactions and the proton donor feature of a
proton attached to the O3 atom of AT (Figure 6).

BO

X Y
O3
Z

Figure 6. Molecular simulation for the interactions between talc and AT molecule. The close distance between ring and
the (010) surface indicated possible π-bonding. Dashed points indicate hydrogen bonding interactions.
Adsorption of Atenolol on Talc 391

The results from this study suggest that interactions between AT and talc do exist in aqueous
suspension. Although the maximum amount of AT uptake was only 11 mmol/kg, or less than 0.2%
by weight, caution should be excised if talc is used as the filter or excipient for AT formulation.

REFERENCES

Alderborn, G. (2013) Tablets and compaction. In: M.E. Aulton and K.M. Taylor, editors. Aulton’s
Pharmaceutics: The Design and Manufacture of Medicines, Elsevier Health Sciences, Philadelphia, PA,
p. 505.
Ashford, M. (2013) Bioavailability—physicochemical and dosage form factors. In: M.E. Aulton and K.M.
Taylor, editors. Aulton’s Pharmaceutics: The Design and Manufacture of Medicines, Elsevier Health
Sciences, Philadelphia, PA, p. 314.
Avdeef, A., Berger, C.M. and Brownell, C. (2000) Pharm. Res. 17, 85.
Barbieri, M., Licha, T., Nödler, K., Carrera, J., Ayora, C. and Sanchez-Vila, X. (2012) Chemosphere. 89, 1272.
Bergaya, F. and Lagaly, G. (2006) General introduction: Clays, clay minerals, and clay science. In:
F. Bergaya, B.K.G. Theng and G. Lagaly, editors. Handbook of Clay Science, Elsevier, Oxford, U.K.,
pp. 1–18.
Bozdağ-Pehlivan, S., Subaşi, B., Vural, I., Unlü, N. and Capan, Y. (2011) Acta Pol. Pharm. 68, 423.
Burdukova, E., Becker, M., Bradshaw, D.J. and Laskowski, J.S. (2007) J. Colloid Interface Sci. 315, 337.
Castro, R.A.E., Canotilho, J., Barbosa, R.M. and Redinha, S.J. (2007a) Spectrochim. Acta, Part A. 67, 1194.
Castro, R.A.E., Canotilho, J., Barbosa, R.M., Silva, A.M., Beja, J.A. and Redinha, S.J. (2007b) Cryst.
Growth Des. 7, 496.
Deng, Y., Wu, F., Liu, B., Hu, X. and Sun, C. (2011) Chem. Eng. J. 174, 571.
Du, H., Yin, X., Ozdemir, O., Liu, J., Wang, X., Zheng, S. and Miller, J.D. (2012) Molecular dynamics
simulation analysis of solutions and surfaces in nonsulfide flotation systems. In: Molecular Modeling for
the Design of Novel Performance Chemicals and Materials, B. Rai, editor. CRC Press, Boca Raton, FL,
p. 108.
Ertl, P., Rohde, B. and Selzer, P. (2000) J. Med. Chem. 43, 3714.
Hansen, S.H., Pedersen-Bjergaard, S. and Rasmussen, K.E. (2011) Introduction to Pharmaceutical Chemical
Analysis, John Wiley & Sons, Chichester, West Sussex.
Hapeshi, E., Achilleos, A., Vasquez, M.I., Michael, C., Xekoukoulotakis, N.P., Mantzavinos, D. and
Kassinos, D. (2010) Water Res. 44, 1737.
Hu, Y., Fitzgerald, N.M., Xing, X., Lv, G., Jiang, W.-T. and Li, Z. (2015) Adv. Mater. Sci. Eng. 2015.
Ibezim, E.C., Ofoefule, S.I., Ejeahalaka, C.N.C. and Orisakwe, O.E. (1999) Am. J. Ther. 6,199.
Jadhav, N.R., Paradkar, A.R., Salunkhe, N.H., Karade, R.S. and Mane, G.G. (2013) World J. Pharm. Pharm.
Sci. 2, 4639.
Kibbey, T.C., Paruchuri, R., Sabatini, D.A. and Chen, L. (2007) Environ. Sci. Technol. 41, 5349.
Kumar, S., Das, M., Gupta, S.K., Kumar, R., Chongdar, A. and Ghosh, L.K. (2013) Pharm. Lett. 5, 436.
Küster, A., Alder, A.C., Escher, B.I., Duis, K., Fenner, K., Garric, J., Hutchinson, T.H., Lapen, D.R., Péry,
A., Römbke, J., Snape, J., Ternes, T., Topp, E., Wehrhan, A. and Knacker, T. (2010) Integr. Environ. Assess.
Manage. 6, 514.
Lv, G., Li, Z., Hoeppner, N., Wu, L. and Liao, L. (2014) Colloids Surf. A. 446, 172.
Manallack, D.T. (2007) Med. Chem. 1, 25.
Niedbala, A., Schaffer, M., Licha, T., Nödler, K., Börnick, H., Ruppert, H. and Worch, E. (2013)
Chemosphere. 90, 1945.
Nikolic, C.V., Nikolic, L.J., Stankovic, M., Kapor, A., Popsavin, M. and Cvetkovic, D. (2007) J. Serb. Chem.
Soc. 72, 737.
Novickis, R.W., Martins, M.V.S., Miranda, L.F., Ribeiro, R.R., Silva, L. and Munhoz, A.H. (2013) Adv. Sci.
Technol. 86, 102.
392 Z. Li et al./Adsorption Science & Technology Vol. 33 No. 4 2015

Pawar, A., Paradkar, A., Kadam, S.A. and Mahadik, K. (2004) AAPS PharmSciTech. 5, 57.
Poudel, B.K. and Thapa, P. (2013) Kathmandu Univ. J. Sci. Eng. Technol. 9, 1.
Rakić, V., Rakić, N., Daković, A. and Auroux, A. (2013) Microporous Mesoporous Mater. 166, 185.
Rao, D.K., Damodharam, T., Yadav, J.S. and Babu, P.S. (2013) Int. J. Pharm. Chem. Sci. 2, 355.
Sable, K.S., Ghadge, N.N., Hole, M.B. and Gadhave, M.V. (2012) Int. J. Inst. Pharm. Life Sci. 2, 2249.
Sancho, J.L.S., Rodríguez, A.R., Torrellas, S.Á. and Rodríguez, J.G. (2012) Desalin. Water Treat. 45, 305.
Schaffer, M., Boxberger, N., Börnick, H., Licha, T. and Worch, E. (2012) Chemosphere. 87, 513.
Seelig, A., Gottschlich, R. and Devant, R.M. (1994) Proc. Natl. Acad. Sci. U S A. 91, 68.
Seema and Datta, M. (2013) Eur. Chem. Bull. 2, 942.
Shirsand, S.B., Swamy, P.V. and Keshavshetti, G.G. (2011) RGUHS J. Pharm. Sci. 1, 4.
Silveira, J.L.R., Dib, S.R. and Faria, A.M. (2014) Anal. Sci. 30, 285.
Spark, D.L. (2003) Environmental Soil Chemistry, 2nd Ed, Academic Press, San Diego, CA, pp. 352.
Sprynskyy, M., Gadzała-Kopciuch, R., Nowak, K. and Buszewski, B. (2012) Colloids Surf., B. 94, 7.
Vilà, J.I., Calpena, A.C., Obach, R. and Domenech, J. (1992) Int. J. Clin. Pharmacol. Ther. Toxicol. 30, 280.
Vinnakota, S.N.A., Deveswaran, R., Bharath, S., Basavaraj, B.V. and Madhavan, V. (2011) Curr. Pharm. Res.
2, 385.
Wang, J. and Somasundaran, P. (2005) J. Colloid Interface Sci. 291, 75.
Weich, A., De Oliveira, D.C., De Melo, J., Goebel, K. and Rolim, C.M.B. (2007) Lat. Am. J. Pharm. 26, 765.
Yamamoto, H., Nakamura, Y., Moriguchi, S., Nakamura, Y., Honda, Y., Tamura, I., Hirata, Y., Hayashi, A.
and Sekizawa, J. (2009) Water Res. 43, 351.

You might also like