+ All documents
Home > Documents > Functionalised pseudo-boehmite nanoparticles as an excellent adsorbent material for anionic dyes

Functionalised pseudo-boehmite nanoparticles as an excellent adsorbent material for anionic dyes

Date post: 19-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
9
Functionalised pseudo-boehmite nanoparticles as an excellent adsorbent material for anionic dyesAnthony R. Auxilio, a Philip C. Andrews, a Peter C. Junk, a Leone Spiccia, * a Daniel Neumann,b Warwick Raverty, b Nafty Vanderhoek b and Jennifer M. Pringle c Received 9th October 2007, Accepted 12th March 2008 First published as an Advance Article on the web 2nd April 2008 DOI: 10.1039/b715545j Pseudo-boehmite has been functionalised with L-lysine by refluxing an aqueous solution containing these two reactants overnight. The resulting nanosized (<10 nm) product is insoluble in water and has been characterised by solid-state NMR spectroscopy, powder X-ray diffraction analysis, N 2 adsorption–desorption analysis and zeta potential measurements. The affinity of this new nanostructured organic–inorganic hybrid material for anionic dyes has been quantified using UV-vis spectrophotometry and by constructing the adsorption isotherms for Acid Blue 9 (AB9), Acid Yellow 23 (AY23), and Acid Red 37 (AR37). Elemental/micro analyses indicate that one lysine molecule is covalently bonded to every 8 nm 2 of the functionalised material giving a composition [(AlOOH) 230 $(H 2 O) 86 $(C 6 N 2 O 2 H 15 )]. The introduction of the positively charged amino groups resulted in a tremendous increase in dye affinity in contrast to the unfunctionalised material. The adsorption isotherms of the functionalised pseudo-boehmite were fitted to the Langmuir model and yielded equilibrium binding constants (K a ) of 2.6 10 3 M 1 for AB9, 1.5 10 5 M 1 for AY23 and 8.4 10 4 M 1 for AR37. AR37 gave a higher monolayer coverage (C m ) value of 0.13 mmol g 1 than AB9 (0.085 mmol g 1 ) and AY23 (0.081 mmol g 1 ). Dye adsorption is correlated with surface coverage of L-lysine and, in the case of AR37, two dye molecules are concluded to be adsorbed per L-lysine while for AY23 a multi-point interaction is proposed to result in a lower dye capacity and a relatively higher affinity of this dye for FPB when compared with AR37. Introduction The development of functionalised inorganic nanoparticles to tailor surface properties continues to be an active area of research. A considerable amount of work has been undertaken to modify the surface of inorganic nanoparticles through non- covalent or electrostatic interactions. 1–5 However, many poten- tial applications 6 require the functional material to be stable enough at moderate temperatures and to retain the potentially useful functionality when processed (chemical/physical), alone or with other materials, to produce products of commercial value. This has prompted us to investigate a surface modification route that will bind (coordinate) the functional moiety strongly to the nanoparticles. When coated on a paper substrate, this new nanomaterial is envisaged to encapsulate anionic dyes during ink-jet printing. A good starting point would be a well known, relatively low cost material which exhibits a high surface area. Pseudo- boehmite is one well-studied aluminium hydroxide having the general formula of AlO(OH)$nH 2 O. In contrast to boehmite, pseudo-boehmite has a smaller particle size and has more crystal defects, leading to broader lines in its X-ray diffraction pattern. 7–10 The aluminium ions are situated in an octahedral position. 11 Wefers and Misra, 10 and Tsukada et al., 12 have reported that water binds more strongly to pseudo-boehmite than the more crystalline boehmite, although this may not affect the lattice parameters. 13 Both of these materials are amphoteric and are more soluble in acidic and alkaline solutions. 14 Pseudo-boehmite has found useful applications as catalyst supports, 15,16 in anti- corrosion, 17,18 deodorants 19,20 and UV-protection agents. 21 Particularly interesting is its wide usage as an adsorbent material in areas ranging from chemical instrumentation 22 to biochemical systems 23,24 and from environmental pollution control 25,26 to paper coatings. 27 Previously, we reported the adsorption and intercalation of Acid Blue 9 on Mg/Al layered double hydroxides, focusing on the effect of varying the metal composition (Mg/Al) on its adsorptive properties. 28 We have now investigated the surface modification of pseudo-boehmite with L-lysine in an effort to produce a water insoluble material that has higher affinity for anionic dyes commonly used in the printing industry. This work can be contrasted to that of Callender 29 who reported the a CRC Smartprint and School of Chemistry, Monash University, Victoria, 3800, Australia. E-mail: [email protected]; Fax: +61 3 9905 4597; Tel: +61 3 9905 4526 b CRC Smartprint and CSIRO Ensis Papro, Clayton, Victoria, 3169, Australia c ARC Centre of Excellence for Electromaterials Science, School of Chemistry, Monash University, Victoria, 3800, Australia † Electronic supplementary information (ESI) available: (1) Thermogravimetric profiles of pseudo-boehmite (PB) and functionalised pseudo-boehmite (FPB); (2) 27 Al MAS NMR spectra of PB and FPB; (3) results of preliminary dye affinity tests; and (4) three-dimensional structure and molecular dimensions of AB9, AY23 and AR37 from Hyperchem simulation. See DOI: 10.1039/b715545j ‡ Present address: Human Protection and Performance Division, DSTO, PO Box 4331, Melbourne 3001, Australia. 2466 | J. Mater. Chem., 2008, 18, 2466–2474 This journal is ª The Royal Society of Chemistry 2008 PAPER www.rsc.org/materials | Journal of Materials Chemistry Published on 02 April 2008. Downloaded by Monash University on 08/11/2013 01:20:59. View Article Online / Journal Homepage / Table of Contents for this issue
Transcript

PAPER www.rsc.org/materials | Journal of Materials Chemistry

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online / Journal Homepage / Table of Contents for this issue

Functionalised pseudo-boehmite nanoparticles as an excellent adsorbentmaterial for anionic dyes†

Anthony R. Auxilio,a Philip C. Andrews,a Peter C. Junk,a Leone Spiccia,*a Daniel Neumann,‡b

Warwick Raverty,b Nafty Vanderhoekb and Jennifer M. Pringlec

Received 9th October 2007, Accepted 12th March 2008

First published as an Advance Article on the web 2nd April 2008

DOI: 10.1039/b715545j

Pseudo-boehmite has been functionalised with L-lysine by refluxing an aqueous solution containing

these two reactants overnight. The resulting nanosized (<10 nm) product is insoluble in water and has

been characterised by solid-state NMR spectroscopy, powder X-ray diffraction analysis, N2

adsorption–desorption analysis and zeta potential measurements. The affinity of this new

nanostructured organic–inorganic hybrid material for anionic dyes has been quantified using UV-vis

spectrophotometry and by constructing the adsorption isotherms for Acid Blue 9 (AB9), Acid Yellow

23 (AY23), and Acid Red 37 (AR37). Elemental/micro analyses indicate that one lysine molecule is

covalently bonded to every 8 nm2 of the functionalised material giving a composition

[(AlOOH)230$(H2O)86$(C6N2O2H15)]. The introduction of the positively charged amino groups

resulted in a tremendous increase in dye affinity in contrast to the unfunctionalised material. The

adsorption isotherms of the functionalised pseudo-boehmite were fitted to the Langmuir model and

yielded equilibrium binding constants (Ka) of 2.6 � 103 M�1 for AB9, 1.5 � 105 M�1 for AY23 and

8.4 � 104 M�1 for AR37. AR37 gave a higher monolayer coverage (Cm) value of 0.13 mmol g�1 than

AB9 (0.085 mmol g�1) and AY23 (0.081 mmol g�1). Dye adsorption is correlated with surface

coverage of L-lysine and, in the case of AR37, two dye molecules are concluded to be adsorbed per

L-lysine while for AY23 a multi-point interaction is proposed to result in a lower dye capacity and

a relatively higher affinity of this dye for FPB when compared with AR37.

Introduction

The development of functionalised inorganic nanoparticles to

tailor surface properties continues to be an active area of

research. A considerable amount of work has been undertaken to

modify the surface of inorganic nanoparticles through non-

covalent or electrostatic interactions.1–5 However, many poten-

tial applications6 require the functional material to be stable

enough at moderate temperatures and to retain the potentially

useful functionality when processed (chemical/physical), alone or

with other materials, to produce products of commercial value.

This has prompted us to investigate a surface modification route

that will bind (coordinate) the functional moiety strongly to the

nanoparticles. When coated on a paper substrate, this new

aCRC Smartprint and School of Chemistry, Monash University, Victoria,3800, Australia. E-mail: [email protected]; Fax: +61 39905 4597; Tel: +61 3 9905 4526bCRC Smartprint and CSIRO Ensis Papro, Clayton, Victoria, 3169,AustraliacARC Centre of Excellence for Electromaterials Science, School ofChemistry, Monash University, Victoria, 3800, Australia

† Electronic supplementary information (ESI) available: (1)Thermogravimetric profiles of pseudo-boehmite (PB) andfunctionalised pseudo-boehmite (FPB); (2) 27Al MAS NMR spectra ofPB and FPB; (3) results of preliminary dye affinity tests; and (4)three-dimensional structure and molecular dimensions of AB9, AY23and AR37 from Hyperchem simulation. See DOI: 10.1039/b715545j

‡ Present address: Human Protection and Performance Division, DSTO,PO Box 4331, Melbourne 3001, Australia.

2466 | J. Mater. Chem., 2008, 18, 2466–2474

nanomaterial is envisaged to encapsulate anionic dyes during

ink-jet printing.

A good starting point would be a well known, relatively low

cost material which exhibits a high surface area. Pseudo-

boehmite is one well-studied aluminium hydroxide having the

general formula of AlO(OH)$nH2O. In contrast to boehmite,

pseudo-boehmite has a smaller particle size and has more crystal

defects, leading to broader lines in its X-ray diffraction pattern.7–10

The aluminium ions are situated in an octahedral position.11

Wefers and Misra,10 and Tsukada et al.,12 have reported that

water binds more strongly to pseudo-boehmite than the more

crystalline boehmite, although this may not affect the lattice

parameters.13 Both of these materials are amphoteric and are

more soluble in acidic and alkaline solutions.14 Pseudo-boehmite

has found useful applications as catalyst supports,15,16 in anti-

corrosion,17,18 deodorants19,20 and UV-protection agents.21

Particularly interesting is its wide usage as an adsorbent material

in areas ranging from chemical instrumentation22 to biochemical

systems23,24 and from environmental pollution control25,26 to

paper coatings.27

Previously, we reported the adsorption and intercalation of

Acid Blue 9 on Mg/Al layered double hydroxides, focusing on

the effect of varying the metal composition (Mg/Al) on its

adsorptive properties.28 We have now investigated the surface

modification of pseudo-boehmite with L-lysine in an effort to

produce a water insoluble material that has higher affinity for

anionic dyes commonly used in the printing industry. This work

can be contrasted to that of Callender29 who reported the

This journal is ª The Royal Society of Chemistry 2008

Fig. 1 Salts of (a) Acid Blue 9; (b) Acid Yellow 23 and (c) Acid Red 37.

Table 1 Elemental analyses of functionalised pseudo-boehmite (FPB)

C% H% N% Al%

Experimental 0.62 2.46 0.18 39.60Calculateda 0.46 2.69 0.18 40.08

a Calculated for [(AlOOH)230$(H2O)86$(C6N2O2H15)].

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

synthesis of water soluble carboxylate–alumoxane bearing

a small organic moiety. Their method was exploited by Vogelson

et al.30 to produce nanometer-sized chemically reactive fillers to

be incorporated into thermoset polymers. Horch et al.31 also

utilized the surface-modified carboxylate alumoxane nano-

particles to produce a composite material for bone tissue

engineering scaffolds.

Our new functionalised organic–inorganic hybrid has been

characterised via solid state 13C and 27Al NMR spectroscopy,

powder X-ray diffraction, elemental/microanalysis, N2 adsorp-

tion–desorption analysis and electron scanning microscopy. Zeta

potential measurements were used to assess the pH dependence

of the surface charge density of the material after decoration with

the amine groups. The affinities of some anionic dyes (Acid Blue

9, Acid Yellow 23, and Acid Red 37; see Fig. 1) to the hybrid

nanomaterial were compared with that of the undecorated

pseudo-boehmite.

Table 2 Some properties of commercial dyes used in this work

NameDyecontent (%)

Molecularmass/g mol�1 3exp

a/M�1 cm�1 lmaxa/nm

AB9 85 792.85 1.37 � 105 629AY23 70 534.36 2.57 � 104 426AR37 90 514.53 1.88 � 104 513

a UV-vis analysis in our study.

Materials and methods

Synthesis

A typical reaction leading to functionalised pseudo-boehmite

nanoparticles involved refluxing overnight 30.5 g of L-lysine

monohydrochloride (Sigma-Aldrich) and 10.0 g of pseudo-

boehmite (Sasol) in 200 mL of de-ionised (Milli-Q) water. All

chemicals were used without further purification. The pH of the

mixture was measured prior to and after refluxing and found to

be the same, i.e., ca. 7.0 (� 0.2). At this pH, lysine will carry

a total positive charge.32 The resulting slurry was filtered and the

This journal is ª The Royal Society of Chemistry 2008

filtrate containing any unreacted lysine was discarded. The white

precipitate was washed once with distilled water by stirring

slowly for about 2 min. Care was taken not to over-expose the

product to water as the product converts to gibbsite forming a gel

which is difficult to separate from the liquid phase. The white

product was air dried at room temperature for at least 36 h. It

was then ground with an agate mortar and pestle prior to

characterisation and used in dye adsorption experiments. C,H,N

and Al analyses of the functionalised pseudo-boehmite (FPB)

material are shown in Table 1.

Dye affinity/adsorption experiments

Dye solutions consisting of 1 mM of Acid Blue 9 (Saujanya

Dyechem), 5 mM of Acid Yellow 23 (Sigma-Aldrich) and 5 mM

of Acid Red 37 (Sigma-Aldrich) were prepared in water (Milli-Q)

from the as-received dyes taking into account the dye content

(Table 2). Table 2 also lists the experimental wavelength of

maximum absorbance (lmax) and molar absorptivity (3). The dye

concentrations were based on commercial ink-jet ink concen-

trations (note higher 3 value of AB9). In the absence of crystal

structures, the gas phase molecular structure and size of the dyes

were estimated using HyperChem software to find the lowest

energy conformation (MM+) prior to geometry optimization

(AM1-UHF).

Commercial pseudo-boehmite (PB) and the functionalised

pseudo-boehmite (FPB) were subjected to dye affinity tests

(triplicates) according to a modification of a published method,28

using Acid Blue 9 (AB9), Acid Yellow 23 (AY23) and Acid Red

37 (AR37) solutions. The dye solution (5 mL) was added to the

pigment (PB or FPB; 0.05 g) in a centrifuge tube. The tube was

agitated using a vortex mixer (2500 rpm) for 1 min followed by

centrifugation (3000 rpm, 10 min), as described in previous

work.28 The supernatant liquid was carefully removed and an

aliquot of 25 mL was diluted with distilled water (5 mL) giving

a ratio of 1 : 200 and was analyzed using a VARIAN CARY 300

Bio UV-vis spectrophotometer. For the FPB system, a different

dilution factor was used so that the absorbance reading was in

the measurable range. Absorbances were recorded at the lmax of

visible absorption (see Fig. 2). The amount of dye adsorbed per

unit area (molecules nm�2) was calculated according to:

J. Mater. Chem., 2008, 18, 2466–2474 | 2467

Fig. 2 UV-vis spectra of untreated anionic dyes. Dye concentrations:

Acid Blue 9 (AB9) ¼ 5 mM; Acid Yellow 23 (AY23) ¼ 25 mM; Acid Red

37 (AR37) ¼ 25 mM.

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

dye adsorbed ¼6:02 � 1023

�mol dyetotal � mol dyesupernatant

0:05 g � surface area � 1018

For the determination of adsorption isotherms, varied amounts

of materials (1.0–0.05 g) were exposed to constant dye concen-

tration. Triplicate analyses were also conducted and the super-

natant liquids were analyzed in the same manner as described

above in the dye affinity test. Analysis of the adsorption

isotherms applied the Langmuir equation33 originally derived for

gas adsorption on planar surfaces,34 Cs ¼ CmKaCp/(1 + KaCp),

where Cs (mmol g�1) is the amount of dye adsorbed to the

surface, Cp (mmol mL�1) is the amount of dye remaining in the

solution, Cm (mmol g�1) is the maximum adsorption capacity at

a monolayer coverage, and Ka (M�1) is the equilibrium binding

constant.

Characterisation

Solid-state NMR spectra were measured using a magic angle

spinning speed of 8 KHz on a Bruker AM300 instrument

equipped with a Bruker 4 mm solid-state probe operating at

75.5 MHz for 13C and 78.2 MHz for 27Al. The 13C signal was

enhanced using cross 1H to 13C polarization techniques. For the13C CP MAS NMR spectra a contact time of 1000 ms was used,

with a 3 ms pulse, a recycle delay of 1 s and no line broadening.

Chemical shifts were referenced to external samples of glycine

and Al2O3. For 27Al NMR spectra, a recycle delay of 2 s was

used, with a 5.5 ms pulse and no line broadening.

Powder X-ray diffraction analyses were conducted using

a Philips 1140 diffractometer under the following conditions: 40

kV, 25 mA, CuKa (l ¼ 0.15406 nm) radiation. The samples, as

unoriented powders, were scanned in steps of 0.02� (2q) in the

range from 2 to 70� at a speed of 2� min�1. The apparent

crystallite size was estimated from the width of the (020) reflec-

tion for PB and FPB using Scherrer’s equation.35 The (100) line

of a powdered quartz external standard was used to measure

instrumental broadening.

Surface morphologies and particle size distributions were

studied using a Philips XL30 Field Emission Scanning Electron

Microscope manufactured by Oxford Instruments at a voltage

of 5 kV.

2468 | J. Mater. Chem., 2008, 18, 2466–2474

Zeta potential measurements (triplicate) were carried out using

a ZetaPALS instrument manufactured by Brookhaven Instru-

ments Corporation fitted with an autotitrator (BI-ZTU) at room

temperature. An aliquot of 300 ppm ultrasonically dispersed

particles containing 1 mM KNO3 electrolyte was placed in a flow

cell. Acidity and basicity were adjusted using 0.1 M or 1 mM

HNO3 or KOH as required.

Surface area, pore volume, pore size, and pore size distribution

measurements were made using a Micromeritics ASAP 2010

instrument. The samples were degassed at 378 K prior to analysis

using nitrogen adsorption–desorption at 77 K.

The density was determined by the helium pycnometry tech-

nique using an AccuPyc 1330 gas displacement pycnometer

manufactured by Micromeritics. Samples were dried at 333 K

overnight in static air and placed in a desiccator prior to analysis.

Microanalysis was performed by the Campbell Microana-

lytical Laboratory, University of Otago, New Zealand and Dairy

Technical Services Ltd, Victoria, Australia. Aluminium content

was determined via Inductively Coupled Plasma–Atomic Emis-

sion Spectroscopy (ICP-AES) by ALS Environmental, Victoria,

Australia.

Results and discussion

Surface modification of pseudo-boehmite

Recent work in our laboratory has focussed on producing an

insoluble inorganic-based material suitable as an inorganic

pigment in the formulation of an ink-receptive layer for paper

coatings. To this end, we have investigated the reaction of

pseudo-boehmite (PB) with L-lysine in water. Modification of PB

involved refluxing PB with L-lysine overnight at pH ca. 7.0. Such

materials do have some solubility over a broad range of pH, but

the rate of dissolution is considerably pH-dependent, with

a minimum near neutral pH.36 Surface modification could

therefore be achieved via either direct reaction of the deproto-

nated carboxylate on L-lysine with the PB surface or by a

dissolution–chemical reaction (also involving carboxylate

binding)–reprecipitation mechanism.

Unequivocal evidence of the carboxylate group of lysine being

coordinated to aluminium atoms of pseudo-boehmite is provided

by 13C CP MAS NMR spectroscopy. The chemical shifts for the

carboxylate and aliphatic carbons are expected to be at 160–190

ppm and 0–62 ppm respectively.37 Fig. 3 shows the 13C CP MAS

NMR spectra of PB, FPB and pure L-lysine monohydrochloride.

The 13C NMR spectrum of pure lysine (Fig. 3a) shows the

carboxylate carbon signal peak at 178 ppm. However, for the

lysine functionalised pseudo-boehmite (Fig. 3b) two resonances

can be observed at 167 and 184 ppm. The latter can be assigned

to the carboxylate carbon of lysine coordinated to alumi-

nium(III). According to Barron et al.,11,38 carboxylate groups can

have either bridging (Fig. 4a) or terminal monodentate ligation

(Fig. 4b) to aluminium and not chelation (Fig. 4c), which is

unfavoured due to ring strain. In our study, the significant

downfield shift of ca. 6 ppm for the carboxylate carbon supports

a bridging interaction since an increase in polarization of the

carboxyl bond is expected for bridging compared with the

terminal monodentate ligation. The second resonance at 167

ppm could be carboxylate or carbonate ion chelating to the

This journal is ª The Royal Society of Chemistry 2008

Fig. 3 13C CP MAS NMR spectra of: (a) pure L-lysine$HCl; (b) func-

tionalised pseudo-boehmite (FPB) after washing with water to remove

excess L-lysine; and (c) pseudo-boehmite prior to functionalisation with

L-lysine.

Fig. 4 Possible binding modes of the carboxylate in L-lysine to the

aluminium centres in pseudo-boehmite: (a) syn–syn bridging; (b) syn

monodentate terminal; and (c) symmetric (asymmetric) chelation.

Fig. 5 X-Ray powder diffractograms of: (a) pseudo-boehmite (PB); (b)

functionalised pseudo-boehmite (FPB); and (c) powdered quartz

(included for crystallite size determination). Note: Reflections index of PB

and quartz were based on Potdar et al.44 and Levien et al.,45 respectively.

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

aluminium ions. Narayanan and Laine39 reported that the

carboxylate carbon of gibbsite formate [Al(O2CH)3] gave

a solution 13C NMR resonance at 165 ppm, which is similar to

the commercial aluminium formate (166 ppm) and indicated that

this corresponds to a chelating formate ligand. This signal could

also be due to some HCO3� or CO3

2� groups that adsorbed on

the aluminium surface. In the 13C NMR spectrum of a uranyl(V)

carbonate complex, a signal at 169 ppm was assigned to free

CO32�.40 Thus, in this study the signal at 167 ppm could be due to

some HCO3� or CO3

2� adsorbed on the aluminium surface as

evident in the 13C spectrum of the PB material measured before

functionalisation (Fig. 3c).

This journal is ª The Royal Society of Chemistry 2008

27Al MAS NMR spectra of PB and FPB (ESI)† showed strong

signals at ca. 4 ppm, which can be assigned to six-coordinate

Al(III) centres. In 27Al NMR spectra, signals in the ranges of 300–

200 ppm, 200–70 ppm, ca. 50 ppm and ca. 0 ppm are normally

assigned to three, four, five and six-coordinate aluminium

centres, respectively.41–43 In this study, surface functionalisation

had no effect on the Al(O/OH) coordination environment.

Fig. 5a and b show that the PXRD patterns of the pseudo-

boehmite material before and after functionalisation are essen-

tially identical. Peaks characteristic of the mineral boehmite

(JCPDS-ICDD: 83-2384) confirm that the pseudo-boehmite core

structure does not undergo transformation during the reaction

with the lysine molecule. The broadened peaks can be attributed

to the colloidal nature of the pseudo-boehmite particles46 and

thus smaller particle size. Application of Scherrer’s equation,35

t ¼ 0.9l/bcosqb, where t is the crystallite size, l is the wavelength

of the radiation used, b is the pure diffraction breadth and qb is

the Bragg diffraction angle, enabled us to estimate that the

crystallite sizes of both PB and FPB were between 5–10 nm. This

value is within the values reported by previous studies.11,13,44,47–49

It has been noted11 that the composition of the hybrid material

is difficult to obtain owing to the variation of the physical

characteristics of the individual PB particles, such as particle size,

surface morphology and the identity of surface groups, such as

Al–O–Al or Al–O(H)–Al. Nevertheless, the number of lysine

molecules on the PB surface can be obtained even though the

loading is very low. Microanalyses (see Table 1) were consistent

with the composition of [(AlOOH)230$(H2O)86$(C6N2O2H15)]

and indicate that one lysine molecule is covalently bound per

8 nm2 of FPB surface. This composition is consistent with ther-

mogravimetric analysis (ESI)† giving about 9% of surface bound

water. The observed %C value is slightly higher than the calcu-

lated value probably due to the presence of small amounts of

HCO3� or CO3

2� (<0.8% by mass) which is consistent with our

solid state NMR results. The amount of lysine, <1% (by weight)

of the hybrid material, could not be observed in the thermo-

gravimetric profile (ESI).† Given that there are 65 mmol of lysine

per gram of FPB and assuming a particle diameter of 5–10 nm

J. Mater. Chem., 2008, 18, 2466–2474 | 2469

Fig. 6 Zeta potential (x) of PB and FPB as a function of pH. The

isoelectric point (IEP) of PB is at pH ¼ 9.26 (�0.02) whilst that of FPB is

at pH ¼ 9.69 (�0.04). Curves were fitted using a polynomial equation.

Table 3 Correlation of the amount of dye adsorbed (number of mole-cules or mmol) to lysine coverage (unit area or mmol) of functionalisedpseudo-boehmite (FPB) nanoparticles. On average there is one lysinemolecule per 8 nm2 of FPB surface

Dye

Adsorbed dye

Molecular box dimension/nm (see ESI†)b

Number of dyemolecules per 8 nm2 FPBa x-axis y-axis z-axis

AB9 1.5 1.03 1.00 1.28AY23 1.4 0.74 0.24 1.69AR37 2.2 0.94 0.33 1.54

a Microanalysis, adsorption isotherm (Cm) and BET surface area.b Hyperchem simulation (shown for easy correlation).

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

(PXRD analysis) there would be in the order of 2–10 lysine

molecules bound per isolated FPB nanoparticle.

The isoelectric points (IEP) of the PB and FPB particles were

determined to further confirm the presence of the positively

charged amine groups on the surface of pseudo-boehmite. The

electrophoretic mobility (m) of the particles was measured at

varying pH and the electrokinetic/zeta (x) potential calculated by

the Smoluchowski equation, x ¼ mh/3, where h and 3 are the

viscosity and permittivity of the solution, respectively. A plot of

x potential versus pH can be used to determine IEP (pH of zero x

potential), the pH at which the average surface charge is zero. If

the IEP is <7 the surface is acidic, if it is >7 the surface is

basic.50,51 A plot of the zeta potential as a function of pH for PB

and FPB (Fig. 6) shows that both materials have a similar zeta

potential of ca. +43 mV below pH 7. At the pH of the adsorption

experiments of 7.8, however, the x potential of FPB is 6 mV

higher than that of PB. This should result in greater adsorption

of the dye to FPB. In addition, strong H-bonding interactions

between the dyes and the positively charged groups on L-lysine

would be expected to lead to a higher affinity.

As previously noted, the dissolution of aluminium oxides is at

its minimum in the neutral pH range. In fact, the amount of

aluminium in the supernatant liquid after the dye affinity tests

was negligible (3–5 ppm). The presence of an additional posi-

tively charged group on the PB surface is indicated by an increase

in IEP from ca. 9.3 for PB to 9.7 for FPB. According to Bruice,52

the pKa of the amino groups in lysine (8.95 and 10.79) would give

rise to an isoelectric point of 9.87. Since this value is very close to

the IEP of FPB, the increase in shift in IEP reflects the presence

of lysine amine groups.

Dye affinity/adsorption studies

The attachment of the L-lysine to the pseudo-boehmite surface

introduces positively charged amino groups that leads to

substantially enhanced dye affinity. Initial experiments carried

out at pH ¼ 7.8 indicated that, in addition to having a high

affinity for FPB, AB9 forms adducts with FPB leading to

colloidal stabilization (as described by the Debye–Huckel equa-

tion).51,53 Since the dye adsorption measurements relied on the

2470 | J. Mater. Chem., 2008, 18, 2466–2474

separation of the supernatant dye from the solid FPB, the

experimental conditions were designed to avoid colloidal stabi-

lization (i.e., low masses of FPB to volume of dye solutions were

used). Such problems were not encountered for AY23 and AR37.

Preliminary tests on both PB and FPB indicated that the

affinity of the dyes for the adsorbent follows the order AR37 >

AY23 > AB9 (see ESI)†. A higher amount of AR37 was adsor-

bed which may be due to the lower number of negative charges

on this dye (2 vs. 3 for the others, see Fig. 1). This means that

more dye is needed to balance the positive surface charge. All

three dyes show substantially greater adsorption on FPB than PB

(by ca. two-fold), as would be expected if the oxide surface had

been decorated with positively charged functional groups.

Quantification of the adsorptive properties of both (PB and FPB)

materials is discussed below. Analysis of the data taking account

of the textural properties of PB and FPB (see Table 5, later)

allowed the determination of dye coverage on the surface, as

summarised below in Table 3.

Fig. 7 shows the adsorption isotherms for AB9, AY23 and

AR37 binding to the FPB surface. Fitting of the data was

approximated using the Langmuir model developed for mono-

layer adsorption on a homogeneous surface. Whilst not ideal,

this enabled us to estimate the maximum adsorption capacity

(Cm) and binding constants (Ka). According to the classification

of Giles et al.54 the isotherms for AY23 and AR37 binding to

FPB can be classified as L-2 type curves that indicate a very

strong intermolecular attraction between the adsorbate and the

adsorbent. This is particularly evident for AR37 and AY23, for

which the initial part of the isotherm is nearly vertical. For

a Langmuirean system, the plateau would represents complete

monolayer coverage from which maximum adsorption capacity

can be taken.28 Although our system is not strictly Langmuirean,

the data provide useful information about dye adsorption. Fig. 8

shows the adsorption isotherms for AB9, AY23 and AR37

binding to the untreated PB surface. This sub-type of L curve

does not have a plateau or an inflection indicating that surface

saturation has not been reached. An attempt to reach the

saturation point by increasing the proportion of dye (smaller

amount of PB) gave the same result but with a larger standard

deviation. None of the well known adsorption–sorption models

such as Langmuir, Freundlich and Temkin could satisfactorily fit

the data. The linear increase in dye adsorption with dye

concentration indicates that the dye has little or no affinity for

the unfunctionalised pseudo-boehmite surface. As the surface

This journal is ª The Royal Society of Chemistry 2008

Fig. 7 Adsorption isotherms for: (a) Acid Red 37; (b) Acid Yellow 23;

and (c) Acid Blue 9 binding to functionalized pseudo-boehmite (FPB)

system at pH ¼ 7.8 and room temperature. The full lines represent fitting

of the data to the Langmuir equation.

Fig. 8 Adsorption isotherms for: (a) Acid Red 37; (b) Acid Yellow 23;

and (c) Acid Blue 9 binding to pseudo-boehmite (PB) system at pH � 8.4

and room temperature. Data could not be fitted with the Langmuir,

Freundlich or Temkin models.

Table 4 Langmuir model parameters for the adsorption of Acid Blue 9(AB9), Acid Yellow 23 (AY23) and Acid Red 37 (AR37) binding tofunctionalised pseudo-boehmite (FPB) at pH ¼ 7.8

Dye Cm/mmol g�1 (�10�1) Ka/M�1 (�103)

AB9 0.85 (�0.03) 2.65 (�0.35)AY23 0.81 (�0.03) 151 (�44)AR37 1.28 (�0.05) 84 (�17)

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

concentration of dye increases, the hydrophobicity of the surface

may increase leading to an increase in adsorption.

The Langmuir model parameters (Cm and Ka) for FPB are

listed in Table 4 and reveal interesting trends. The FPB material

has the highest affinity for AY23 (Ka ¼ 1.5 � 105 M�1) followed

by AR37 (8.4 � 104 M�1) and AB9 (2.6 � 103 M�1). These

differences may be correlated with the charge on the dyes (three

for AY23 and two for AR37 and AB9). In addition, positive

charge on AB9 appears to reduce its affinity for FPB. The Ka

value for AY23 is only slightly higher than that for AR37, despite

the fact that in AY23 one of the three charged groups is

a carboxylate which is more basic than the sulfonates and would

be expected to have higher affinity for the boehmite surface. A

likely explanation, which is in keeping with the lower binding

capacity of AY23 (vide infra), is that multipoint interaction is

possible for this dye.

Although we have been unable to determine the parameters

(Cm and Ka) for dye binding to untreated PB surfaces, inspection

of the data in Fig. 7 and 8 indicates that the dye binding capacity

of the modified FPB is over two-fold higher than the untreated

This journal is ª The Royal Society of Chemistry 2008

PB. For FPB, AR37 gave a higher dye coverage (Cm ¼ 0.13

mmol g�1) than AB9 and AY23 (Cm ¼ 0.08 mmol g�1). This

finding is discussed further below.

Computational chemistry was employed to provide some

insight into the three-dimensional shape and size of these dyes in

the gas phase (Table 3; ESI).† The conclusion reached was that

the L-lysine coverage is such that the steric bulk/size of the dye

will not affect dye binding. Table 3 shows the number of dye

molecules adsorbed per 8 nm2 of FPB. If it were assumed that the

binding of the dye results in charge neutralization then every

J. Mater. Chem., 2008, 18, 2466–2474 | 2471

Table 5 Textural properties of pseudo-boehmite (PB) and functional-ised pseudo-boehmite (FPB)

MaterialSurfaceareaa/m2 g�1

Ave. porediameterb/A

Correctedpore vol.c/cm3 g�1

Ave.densityd/g cm�3

PB 284 49 0.35 2.70 (�0.00)FPB 287 44 0.33 2.73 (�0.01)

a BET method. b BJH method. c BJH method. d Helium density.

Fig. 10 N2 Adsorption–desorption isotherms of (a) pseudo-boehmite

and (b) functionalized pseudo-boehmite. Adsorption is represented by

the solid line and desorption by the dashed line.

Fig. 9 Schematic representation of anionic dyes binding to the func-

tionalised pseudo-boehmite (FPB) surface.

2472 | J. Mater. Chem., 2008, 18, 2466–2474

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

lysine molecule (8 nm2 of FPB) requires 1.0 AR37 and 1.0 AB9

(2� charge), and 0.67 AY23 (3� charge). The amount of dye

adsorbed is well above these values (Table 3). One possible origin

of this difference is the adsorption of the dye directly onto the

‘lysine free’ pseudo-boehmite surface (see Fig. 9). Further

explanations for the observed behaviour can be based on the idea

that complete charge neutralization (dye�/L-lysine+) does not

occur, either because the interaction does not involve all anionic

groups on the dye or because of dye aggregation occurs on the

FPB surface. For AR37, the fact that approximately two mole-

cules of dye are adsorbed per molecule of L-lysine indicates that

two dye molecules can access and interact with each ammonium

group on the diprotonated L-lysine (SO3�/+H3N–).

Table 5 summarizes the textural properties of the PB and FPB

particles. The surface area of pseudo-boehmite was found to be

284 m2 g�1, which is in agreement with literature value.55,56 There

is no significant difference between PB and FPB in terms of their

textural properties. This means that the excellent dye adsorptive

behaviour exhibited by the FPB system could only be due to the

presence of the positively charged amine group pendant to the

surface. Technically, these materials are considered to be

mesoporous, i.e., having pore openings between 20 A and 500

A.28 Fig. 10 shows the N2 adsorption–desorption isotherms of PB

and FPB nanoparticles. Clearly, both systems are considered to

be mesoporous because they exhibit a type IV isotherm. Since the

actual material is amorphous in nature, it is likely to have a wide

range of pore sizes.57,58 Regardless of their pore structure, both

the external and internal surfaces will be available for adsorption

as long as they are accessible to the dye adsorbate. The PB and

Fig. 11 Scanning electron micrographs of (a) pseudo-boehmite (PB) and

(b) functionalized pseudo-boehmite (FPB).

This journal is ª The Royal Society of Chemistry 2008

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

FPB surface at which adsorption takes place is assumed to be the

(0, k, 0) plane. It should be pointed out that these isotherms could

hardly be considered to be of type II as classified by Fukasawa

et al.59 since one can observe a limiting adsorption capacity at

high relative pressures. However, with its characteristic hysteresis

loop (H-2 type), Sing et al.57 have noted that although these

materials are definitely porous, the distribution of pore size and

shape is not well defined and is difficult to interpret.

An SEM analysis was also conducted on PB and FPB to

determine the approximate shape and size of the material and to

correlate this physical property to its adsorptive behaviour (see

Fig. 11). The smaller particles referred to in the two materials

have an approximate size of ca. 40–100 nm. To explain the

significant difference of the particle size between the values

reported by PXRD and SEM, we proposed that the particles

observed in the SEM micrograph could be an agglomeration of

crystallites of smaller size. Furthermore, both materials assumed

a globular shape although some irregular shapes are also

apparent. However, the surface morphology does not affect the

adsorption property of these materials to anionic dyes.

Conclusions

The present study demonstrates the facile synthesis of an

organic–inorganic hybrid nanomaterial that is an effective

anionic dye receptor in an aqueous medium can be achieved

through the reaction of pseudo-boehmite with L-lysine. The

presence of the positively charged organic moiety on the surface

of pseudo-boehmite nanoparticles resulted in a substantial

increase in the ability of the pseudo-boehmite material to adsorb

species of opposite charge. This excellent adsorption behaviour

can be correlated to an increase in electrostatic attraction and/or

hydrogen bonding interactions between the positively charged

pseudo-boehmite nanoparticles and the negatively charged

anionic dyes. For AR37, dye adsorption correlated with surface

coverage of L-lysine. That is, two dye molecules were adsorbed

per L-lysine. For AY23, a multi-point interaction is proposed to

account for the lower dye capacity and the high affinity of this

dye for FPB when compared with AR37.

Acknowledgements

This research was supported by the Smartprint Cooperative

Research Centre (CRC), Australia. The authors would like to

acknowledge the assistance that Dr Ian McKinnon provided

with the zeta potential measurements. A.R.A. acknowledges

support provided by Monash University in the form of a Post-

graduate Publication Award.

References

1 F. Caruso, H. Lichtenfeld, E. Donath and H. Mohwad,Macromolecules, 1999, 32, 2317.

2 D. I. Gittins and F. Caruso, Adv. Mater., 2000, 12, 1947.3 D. G. Kurth, F. Caruso and C. Schuler, Chem. Commun., 1999, 1579.4 I. L. Radtchenko, G. B. Sukhorukov, S. Leporatti, G. B. Khomutov,

E. Donath and H. Mohwald, J. Colloid Interface Sci., 2000, 230, 272.5 G. B. Sukhorukov, E. Donath, S. Davis, H. Lichtenfeld, F. Caruso,

V. I. Popov and H. Mohwald, Polym. Adv. Technol., 1998, 9, 759.6 C. Sanchez, B. Julian, P. Belleville and M. Popall, J. Mater. Chem.,

2005, 15, 3559.7 R. Tettenhorst and D. A. Hofmann, Clays ClayMiner., 1980, 28, 373.

This journal is ª The Royal Society of Chemistry 2008

8 M. Digne, P. Sautet, P. Raybaud, H. Toulhoat and E. Artacho,J. Phys. Chem. B, 2002, 106, 5155.

9 M. Nguefack, A. F. Popa, S. Rossignol and C. Kappenstein,Phys. Chem. Chem. Phys., 2003, 5, 4279.

10 K. Wefers and C. Misra, Oxides and Hydroxides of Aluminum, 2ndedn, Alcoa Laboratories, USA, 1987.

11 C. C. Landry, N. Pappe, M. R. Mason, A. W. Apblett, A. N. Tyler,A. N. Macinnes and A. R. Barron, J. Mater. Chem., 1995, 5, 331.

12 T. Tsukada, H. Segawa, A. Yasumori and K. Okada, J. Mater.Chem., 1999, 9, 549.

13 M. L. Guzman-Castillo, X. Bokhimi, A. Toledo-Antonio,J. Salmones-Blasquez and F. Hernandez-Beltran, J. Phys. Chem. B,2001, 105, 2099.

14 S. Castet, J.-L. Dandurand, J. Schott and R. Gout, Geochim.Cosmochim. Acta, 1993, 57, 4869.

15 L. L. Murrell and N. C. Dispenziere, Jr., US Pat., 40 831 007, 1989.16 D. H. Han, O. O. Park and Y. G. Kim, Appl. Catal. B: Gen., 1992, 86,

71.17 S. I. Seok, J. H. Kim, K. H. Choi and Y. Y. Hwang, Surf. Coat.

Technol., 2006, 200, 3468.18 D. G. Altenpohl, Corrosion, 1962, 18, 143t.19 C. Kropf, T. Foerster, M. Heller, M. Claas and B. Banawski, Ger.

Pat., 20 000 615, 2000.20 S. Suzuki, Jpn. Pat., 07 080 289, 1995.21 S. Ishida, N. Takeuchi, K. Fujiyoshi, Y. Michiie, T. Susuki, K. Kido,

H. Kigata and H. Mitsunaka, Jpn. Pat., 20 041 216, 2004.22 J. J. Kirkland, Anal. Chem., 1963, 35, 1295.23 M. Nishida, Y. Yoshimura, J. Kawada, A. Ookubo, T. Kagawa,

A. Ikawa, Y. Hashimura and T. Suzuki, Biochem. Int., 1990, 22, 913.24 R. J. Sepelyak, J. R. Feldkamp, T. E. Moody, J. L. White and

S. L. Hem, J. Pharm. Sci., 1984, 73, 1514.25 M. Kabayama, N. Kawasaki, T. Nakamura, T. Tokimoto and

S. Tanada, Hyomen Kagaku, 2004, 25, 499.26 R. Trejo-Vazquez and A. Bonilla-Petriciolet, Afinidad, 2004, 61, 116.27 J. Shi, T. Schuman and J. Stoffer, JCT Res., 2004, 1, 1.28 A. R. Auxilio, P. C. Andrews, P. C. Junk, L. Spiccia, D. Neumann,

W. Raverty and N. Vanderhoek, Polyhedron, 2007, 26, 3479.29 R. L. Callender, C. J. Harlan, N. M. Shapiro, C. D. Jones,

D. L. Callahan, M. R. Wiesner, D. B. MacQueen, R. Cook andA. R. Barron, Chem. Mater., 1997, 9, 2418.

30 C. T. Vogelson, Y. Koide, L. B. Alemany and A. R. Barron, Chem.Mater., 2000, 12, 795.

31 R. A. Horch, N. Shahid, A. S. Mistry, M. D. Timmer, A. G. Mikosand A. R. Barron, Biomacromolecules, 2004, 5, 1990.

32 N. N. Vlasova and L. P. Golovkova, Colloid J., 2004, 66, 657.33 M. A. Ulibarri, I. Pavlovic, C. Barriga, M. C. Hermosin and

J. Cornejo, Appl. Clay Sci., 2001, 18, 17.34 I. Langmuir, J. Am. Chem. Soc., 1918, 40, 1361.35 B. D. Cullity and S. R. Stock, Elements of X-ray Diffraction, 3rd edn,

Pearson, Prentice Hall, Upper Saddle River, New Jersey, 2001.36 S. M. Kraemer, V. Q. Chiu and J. G. Hering, Environ. Sci. Technol.,

1981, 32, 2876.37 K. S. Rohr and H. W. Spiess, Multidimensional Solid-State NMR and

Polymers, Academic Press, London, 1994.38 J. T. Leman, J. Braddock-Wilking, A. J. Coolong and A. R. Barron,

Inorg. Chem., 1993, 32, 4324.39 R. Narayanan and R. M. Laine, J. Mater. Chem., 2000, 10, 2097.40 K. Mizuoko, I. Grenthe and Y. Ikeda, Inorg. Chem., 2005, 44, 4472.41 A. R. Barron, Polyhedron, 1995, 14, 3197.42 G. Paglia, C. E. Buckley, A. L. Rohl, B. A. Hunter, R. D. Hart,

J. V. Hanna and L. T. Byrne, Phys. Rev. B: Condens. MatterMater. Phys., 2003, 68, 1441.

43 X. Shi, J. Yang and Q. Yang, Eur. J. Inorg. Chem., 2006, 1936.44 H. S. Potdar, K.-W. Jun, J. W. Bae, S.-M. Kim and Y.-J. Lee,

Appl. Catal., A, 2007, 321, 109.45 L. Levien, C. T. Prewitt and D. J. Weidner, Am. Mineral., 1980, 65,

920.46 R. Aucejo, J. Alarcon, C. Soriano, M. C. Guillem, E. G. Espana and

F. Torres, J. Mater. Chem., 2005, 15, 2920.47 X. Bokhimi, J. A. Toledo-Antonio, M. L. Guzman-Castillo and

F. Hernandez-Beltran, J. Solid State Chem., 2001, 159, 32.48 K. Okada, T. Nagashima, Y. Kameshima and A. Yasumori,

J. Colloid Interface Sci., 2002, 248, 111.49 D. Mishra, S. Anand, R. K. Panda and R. P. Das, Mater. Lett., 2000,

42, 38.

J. Mater. Chem., 2008, 18, 2466–2474 | 2473

Publ

ishe

d on

02

Apr

il 20

08. D

ownl

oade

d by

Mon

ash

Uni

vers

ity o

n 08

/11/

2013

01:

20:5

9.

View Article Online

50 Instruction Manual for Using the BI-ZTU - The Autotitrator for Usewith Brookhaven’s Zeta Potential Instruments, BrookhavenInstruments Corp., New York, 2005.

51 A. W. Adamson and A. P. Gast, Physical Chemistry of Surfaces, 6thedn, John Wiley & Sons, Inc., New York, 1997.

52 P. Y. Bruice, Organic Chemistry, 3rd edn, Prentice-Hall Inc.,New Jersey, 2001.

53 D. I. Gittins and F. Caruso, J. Phys. Chem. B, 2001, 105, 6846.54 C. H. Giles, T. H. MacEwan, S. N. Nakhwa and D. Smith, J. Chem.

Soc., 1960, 3973.

2474 | J. Mater. Chem., 2008, 18, 2466–2474

55 J.-F. Hochepied, O. Ilioukhina and M.-H. Berger, Mater. Lett., 2003,57, 2817.

56 P. Cambier and G. Sposito, Clays Clay Miner., 1991, 39, 369.57 K. S. W. Sing, D. H. Everett, R. A. W. Haul, L. Moscou,

R. A. Pierotti, J. Rouquerol and T. Siemieniewska, Pure Appl.Chem., 1985, 57, 603.

58 P. A. Webb and C. Orr, Analytical Methods in Fine ParticleTechnology, Micromeritics Instrument Corp., USA, 1997.

59 J.-i. Fukasawa, H. Tsutsumi, M. Sato and K. Kaneko, Langmuir,1984, 10, 2718.

This journal is ª The Royal Society of Chemistry 2008


Recommended