Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

CSIRO PUBLISHING Full Paper

www.publish.csiro.au/journals/ajc Aust. J. Chem. 2010, 63, 83–91

Adsorption of Ink-Jet Inks and Anionic Dyes onto Mg-Al-NO3


Layered Double Hydroxides of Variable Mg:Al Molar Ratio

Anthony R. Auxilio,A Philip C. Andrews,A Peter C. Junk,A and Leone SpicciaA,B


A School of Chemistry, Monash University, Clayton, Vic. 3800, Australia.
B Corresponding author. Email: leone.spiccia@sci.monash.edu.au

Layered double hydroxides are materials that show promise as adsorbing media for ink-jet printing. In this work, layered
double hydroxides with nitrate as interlayer anions and with variable Mg:Al molar ratios ranging from 2.2 to 8.1 have been
synthesized at constant pH by the coprecipitation method and characterized by powder X-ray diffraction, N2 adsorption–
desorption, scanning electron microscopy, and helium pycnometry techniques. The ability of these materials to adsorb
ink-jet inks (BCI-Cyan, BCI-Yellow, and BCI-Magenta) and anionic dyes (CI Acid Blue 9 (AB9), CI Acid Yellow 23
(AY23), and CI Acid Red 37 (AR37)) were investigated. BCI-Yellow was found to have high affinity for material with
an Mg:Al molar ratio between 2.2 and 4.0, whereas BCI-Cyan and BCI-Magenta showed a high affinity when the Mg:Al
molar ratio was between 4.8 and 5.5. The adsorption isotherms for AB9, AY23, and AR37 on the material with Mg:Al
molar ratios 2.2–5.5 exhibited H-type curves, indicating a very strong interaction between the adsorbate and adsorbent.
Although there was no clear systematic correlation between the Mg:Al molar ratio and the Cm (dye capacity) and Ka (dye
affinity), the optimum Mg:Al molar ratio was 4.0, and this material has the potential to be used in Ink-Receptive Layer
formulation.

Manuscript received: 26 June 2009.


Manuscript accepted: 13 August 2009.

Introduction respectively, for Al/(Mg + Al) values between ∼0.20 and 0.33
A variety of anionic dyes are commonly used in the formulation (Mg:Al = 2–4).[9] LDHs are very well known for their relatively
of aqueous-based ink-jet inks.[1–8] Examples include Acid Blue high adsorption capacities for anionic species. This is in part due
9 (AB9), Acid Yellow 23 (AY23), and Acid Red 37 (AR37) (see to their relatively high surface area, positively charged surface,
Fig. 1). These dyes contain sulfonate and carboxylate groups and high isoelectric point, and intercalation ability.
are generally used as a salt rather than in their acidic forms.[9] There have been few reports dealing with how variations
They are capable of interacting with a positively charged surface in M2+ :M3+ molar ratios in LDHs with different interlayer
through weak (e.g. Van der Waals) and strong (e.g. electrostatic oxo-anions influence the ability of the material to adsorb guest
and H-bonding) interactions.[10] However, when printing onto anionic species into the intra- and inter-layers of LDHs. Ulibarri
a paper surface, the ink-jet dye–substrate interaction needs to and Hermosin[16] found that the MII /MIII ratio (layer charge)
be further improved to obtain high optical density, high resolu- and the nature of the trivalent cation (Al3+ or Fe3+ ) had no
tion, low image print-through to the back side of the paper and effect on 2,4,6-trinitrophenol adsorption by anionic exchange
high resistance of the image to smearing when wet.[11,12] Con- on the Mg-Al-CO3 system but it did for the Mg-Al-Cl system.
sequently, there is continued interest in the development of solid Studies of the sorption of radioactive iodide from aqueous solu-
phases that improve dye binding properties. The Layered Double tions on Mg-Al-NO3 found that 131 I− was sorbed by anionic
Hydroxides (LDHs) are one family of materials currently being exchange in the interlayer of hydrotalcite with an Al/(Mg + Al)
investigated for this purpose. ratio of 0.20 and that 131 I− sorption increased more than three
LDHs are clays consisting of positively charged metal times for higher ratios (lower Mg:Al values).[17] Pristine Zn-Al-
hydroxide layers with interlamellar anions for charge NO3 -LDHs with various Zn:Al ratios (2, 2.6, and 3) have been
neutrality.[13] The general formula is: [M2+ 3+
1−x Mx (OH)2 ]
x+
intercalated with 5-aminosalicylic acid using direct coprecipita-
[(A )x/n ·mH2 O], where M
n− 2+ and M 3+ are divalent and tion and indirect ion-exchange methods.[18] For both methods,
trivalent cations respectively, x is the trivalent metal ratio, higher intercalate loadings were found at a lower Zn:Al ratio
M3+ /(M2+ + M3+ ), that determines layer charge density, and (owing to higher x values) together with a more ordered crystal
An− is an exchangeable anion.[14] One family of anions are structure. The sorption capability of LDH for TcO− 4 could be
− −
the oxo-anions that include CO2− 2−
3 , NO3 , SO4 , OH , CrO4 ,
2−
greatly increased by manipulating the metal cations (M2+ – Mg,
2− 2− 3−
WO4 , S2 O3 , PO4 , and others. [15] LDHs with these anions Zn, Co, Ni, Cu, Fe, Cd, Pd, and Mn; M3+ – Al, Fe, Cr, V, Y, Yb,

generally have characteristic basal spacing that can be deter- La, Ga, and Bi), interlayer anions (CO2− 2−
3 , NO3 , and SO4 ), and

mined directly from powder X-ray diffraction (XRD) data. For M :M molar ratios (1 to 10). The Ni-Al-(NO3 or CO2−
2+ 3+ [19]
3 )
example, LDH-carbonates can be distinguished from LDH- with Ni:Al ratio of 3:1 was found to exhibit the highest sorp-
nitrates as they have basal spacings of ∼7.7 and 8.7 Å, tion capability among all the materials tested and this correlates

© CSIRO 2010 10.1071/CH09362 0004-9425/10/010083


84 A. R. Auxilio et al.

(a) N

(a)
SO
3
SO

Transmittance [%]
3 2Na

(b)
N

(c)
SO3
(d)
(b) CO
2 400 900 1400 1900 2400 2900 3400
Wavenumber (1/cm)
N
N N
N Fig. 2. Fourier-transform (FT)-IR spectra of (a) Mg(OH)2 and Mg-Al-
LDH-NO3 at Mg:Al molar ratio of: (b) 8.1; (c) 5.5; and (d) 2.2.
3Na SO3
O OH
3S
the commercial and functionalized pseudo-boehmite.[24] Recent
H work on the effect of Mg/Al variation on adsorption properties
(c) O N SO of Mg/Al layered double hydroxides with NO−
3
NH2 3 as counter-
ions, herein referred to as Mg-Al-LDH-NO3 , showed that NO− 3
N
N can be more easily replaced with other anions than multiply
charged anions compared with CO2− 3 .
[25–29] With this in mind,
HO
2NH4 we have extended our work on novel high-performance inor-
ganic pigments for use as ink-receptive layer formulations to
the study of the dye adsorption behaviour of Mg-Al-LDH-NO3
SO materials. Textural properties were obtained using powder X-
3
ray diffraction (PXRD), scanning electron microscopy (SEM),
Fig. 1. Structures of: (a) CI Acid Blue 9; (b) CI Acid Yellow 23; and Fourier-transform (FT)-IR, N2 adsorption–desorption, and He
(c) CI Acid Red 37. pycnometry analyses, and used to interpret the dye adsorption
properties.
with the d(003) basal spacing. An increase in the anion sorption
Results and Discussion
capacity of more than two orders of magnitude accompanied
a small increase (∼0.5 Å) in the basal spacing. The adsorption Materials Synthesis and Characterization
of 2,4-dichlorophenoxyacetate on Mg-Al-NO3 LDHs with dif- The synthesis of Mg-Al-LDH-NO3 materials was achieved by
ferent Mg:Al molar ratios (3.3, 2.6, and 2.1) has been studied a coprecipitation method, as described in the Experimental
paying particular attention to the effect of the orientation of the section. The Mg to Al ratio was varied from ∼2.2 to 8.1.
interlayer nitrate.[20] The material with an Al3+ /(Al3+ + Mg2+ ) The materials were subjected to detailed characterization as
molar ratio of 2.1 was found to exhibit low adsorption capacity discussed below.
owing to the parallel orientation of the interlayer nitrate with The IR spectra of Mg-Al-LDH-NO3 with Mg:Al molar ratios
respect to the hydroxide sheet (small basal spacing) and thus of 8.1, 5.5, and 2.2 and of pure Mg(OH)2 (Fig. 2) exhibit
allowing only external adsorption, whereas the material with absorptions at 836, 1384, and 679 cm−1 due to the out-of-plane
a molar ratio of 3.3 (higher charge density) exhibited higher symmetric deformation mode, antisymmetric stretching mode,
adsorption because nitrate adopted a perpendicular orientation and antisymmetric deformation mode of NO− 3 , respectively.
relative to the hydroxide sheet (external and interlayer adsorp- The broad absorption centred at 3466 cm−1 can be ascribed to
tion). Vreysen and Maes[21] investigated the adsorption of humic O–H stretching vibrations due to interlayer H2 O. The forma-
and fulvic acids onto NO− − 2−
3 , Cl , and CO3 intercalated LDHs tion of magnesium hydroxide and a concomitant reduction in
with different Mg:Al ratios (85/15 to 60/40) and found that the the amount of Mg-Al-LDH-NO3 is also apparent as the Mg:Al
Mg:Al ratio was one factor that influenced the adsorption of metal molar ratio increases. The first evidence is the band at
humic substances. A study of the effect of Mg:Al ratio on borate 442 cm−1 , which is due to an Al–O vibration that is intense
and silicate exchange with nitrate interlayer anions found that when Mg:Al = 2.2 and becomes less intense as the Mg:Al
materials with tilted NO− 3 exchanged borate and silicate, whereas molar ratio increases. This absorption was not observed in pure
those with flat NO− 3 did not.
[22] Thus, the adsorption behaviour Mg(OH)2 . The second evidence is the OH stretch at 3696 cm−1
can be tuned by varying the composition of the LDH material. that becomes apparent at Mg:Al = 6.7 and is due to the O–H
We recently reported the adsorption and intercalation of AB9 vibration of the Mg(OH)2 mineral.[30]
on synthesized and calcined Mg-Al LDHs with CO2− 3 as coun- Powder X-ray diffractograms of the Mg-Al-LDH-NO3 mate-
terions with a focus on the effect of Mg:Al variation on its rials with different Mg:Al values are shown in Fig. 3. As

adsorptive properties.[14,23] Optimum adsorption for the syn- mentioned earlier, LDHs with CO2− 3 and NO3 as interlayer
thesized LDH was in the Mg:Al range of 3.1 to 4.4, while it anions exhibit an average d(003) -spacing of 7.7 and 8.7 Å respec-
was at Mg:Al ∼5.6 for the calcined counterparts. We have also tively, when x (layer charge) is ∼0.25 (Mg:Al = 3.0). It is
studied the adsorption behaviour of AB9, AY23, and AR37 on clear that the pure phase of Mg-Al-LDH-NO3 has a larger
Adsorption of Ink-Jet Inks and Anionic Dyes 85

order for materials to behave as excellent adsorbents. As soon as


* another phase begins to evolve (magnesium hydroxide for Mg-
*
Al-LDH-NO3 and hydromagnesite for Mg-Al-LDH-CO3 [14] ),
* Mg(OH) 2
the adsorptive properties decline. The formation of Mg(OH)2 as
the Mg:Al ratio increases can be explained to simply involve the
precipitation of Mg2+ + 2OH− = Mg(OH)2 .[14] There is no sub-
Intensity [a.u.]

Ratio  8.1
Ratio  7.5
sequent formation of hydromagnesite in this case because of the
Ratio  6.7
CO2 -free environment during the synthesis. Note that, although
Ratio  6.0

Ratio  5.5
the side products (magnesium hydroxide or hydromagnesite) of
either hydrotalcite possess a very basic character, they do not
Ratio  4.8 improve the adsorption behaviour of the synthesized material
Ratio  4.0 (see Dye Adsorption Studies).
Ratio  3.1
The physisorption isotherms (see Accessory Publication,
(003)

(006)

Ratio  2.2 Fig. S1) for all Mg-Al-LDH-NO3 materials with variable Mg:Al
molar ratios were obtained via N2 adsorption–desorption analy-
2 12 22 32 42 52 62
sis. The manifestation of hysteresis loops (H3 type in this case)
2θ [] on all isotherms qualifies these isotherms to be classified as type
IV.[32] Therefore, in this respect, Mg-Al-LDH-NO3 materials are
Fig. 3. X-ray powder diffractograms of Mg-Al-LDH-NO3 materials with
different Mg:Al values and for pure magnesium hydroxide (top trace: most
similar to Mg-Al-LDH-CO3 [14] and its calcined products.[23]
intense reflections highlighted by an asterisk). The hysteresis loop is associated with capillary condensation
taking place in mesopores and the limiting uptake over a range
of high relative pressure (p/p0 ). Aggregations of plate-like par-
basal spacing than Mg-Al-LDH-CO3 (see Accessory Publica- ticles that can lead to slit-shaped pores are then predicted. Note
tion, Table S1). The reason for this is that the interaction of that for all materials except those with Mg:Al = 2.2 and 3.1, no
nitrate (single charge) with the brucite-like sheets of the LDH ‘lower closure point’ was observed owing to the steep region
is weaker than for carbonates (double charge). This results in of the desorption branch expected at ∼p/p0 = 0.42 for nitrogen
less compact lamellar sheets and larger basal spacing. As will be at its boiling point. This is referred to as low-pressure hystere-
noted later, Mg-Al-LDH-NO3 exhibited good adsorption prop- sis, which extends to the lowest attainable pressures. According
erties within a range (2.2–5.5) of Mg:Al molar ratios. For ratios to Sing et al.,[32] this phenomenon may be associated with the
above Mg:Al = 5.5, this type of material exhibited poor affinity swelling of a non-rigid porous structure or with an irreversible
for all dyes studied. Powder X-ray analysis also revealed that an chemical interaction of the adsorbate with the adsorbent. As
Mg:Al = 5.5 ratio (x = 0.15) seems to be the upper limit of the noted before, the formation of Mg(OH)2 as the Mg:Al molar ratio
divalent to trivalent metal ratio for producing a pure phase of increases could not be avoided. The N2 adsorption–desorption
this hydrotalcite-like compound. Above this molar ratio, a new isotherm of Mg(OH)2 (see Accessory Publication, Fig. S1) is
phase can be observed as revealed by the emergence of a new set also type IV with an H3-type hysteresis.
of reflections in the diffraction pattern, notably at 2θ values of The surface area, pore diameter, pore volume and den-
18.6◦ , 38.0◦ , and 50.8◦ (shown in asterisks).These reflections are sity of the Mg-Al-LDH-NO3 materials with different Mg:Al
due to Mg(OH)2 (brucite; JCPDS no: 44–1482) that inevitably molar ratios as determined using the Brunauer-Emmett-Teller
forms as the amount of added magnesium exceeds that required (BET) theory, Barrett-Joyner-Halenda (BJH) method and He
to generate the LDH phase.The trend of the d-spacing for Mg-Al- pycnometry are summarized in Table 1. In general, the
LDH-NO3 decreases as the Mg:Al molar ratio increases (lower surface areas, densities and pore volumes of Mg-Al-LDH-
x) within a range of Mg:Al molar ratio values, starting from the NO3 are similar to Mg-Al-LDH-CO3 reported in a previous
highest value of 8.9 Å for Mg:Al = 2.2 to a lowest value of 8.0 Å study,[14] but significantly lower than the calcined Mg-Al-
for Mg:Al = 4.0. However, at Mg:Al = 4.8 and above, the d- LDH-CO3 .[23] With respect to pore size, the trend is: Mg-Al-
spacing is constant (average 8.2 Å) (see Accessory Publication, LDH-CO3 > calcined Mg-Al-LDH-CO3 > Mg-Al-LDH-NO3 .
Table S1). In comparison with Mg-Al-LDH-CO3 , the d-spacing Nevertheless, these materials are mesoporous (pore diameter =
at very low x is ∼8.1 Å. The less compact textural properties 20–500 Å). Although more Mg(OH)2 forms as the Mg:Al ratio
that partly accounted for the AB9 affinity to Mg-Al-LDH- increases, this does not appear to be reflected in the properties
CO3 in our previous work,[14] cannot explain the adsorption derived from N2 adsorption–desorption.
behaviour of Mg-Al-LDH-NO3 . Although the reasoning may The similarity in the physisorption data for the Mg-Al-LDH-
be valid for AY23, it cannot explain the very low affinity of AB9 NO3 materials studied with Mg-Al-LDH-CO3 [14] having a type
and AR37 for the Mg-Al-LDH-NO3 material with the lowest IV isotherm with an H3-type hysteresis that is also similar to the
Mg:Al value (2.2), the material with the highest d(001) spac- calcined products of Mg-Al-LDH-CO3 [23] suggests that these
ing. Previous studies[28,30] have indicated that the higher the materials have the same particle morphology. In fact, scanning
layer charge (lower Mg:Al), the higher the adsorption capacity electron micrographs (Fig. 4) of Mg-Al-LDH-NO3 with differ-
because the nitrates can be vertically oriented and thus interca- ent Mg:Al molar ratios reveal the same general morphology –
lation can easily occur. This cannot account for the adsorption hexagonal platelets. Other general observations can be made,
behaviour of all dyes used in the present study. The mecha- such as when Mg:Al ≥ 6.0, a more rounded-shape particle can
nism involved in the adsorption process is more complex, as be observed in addition to more detectable large agglomerates,
demonstrated in recent work by Marangoni et al.,[31] which found which is more apparent at Mg:Al = 8.1. The morphology for
variations in the adsorption of different bulky anionic dyes into the Mg:Al = 4.8 material is noteworthy as it looks different
zinc hydroxide nitrate. Nevertheless, the current work highlights compared with the rest of the materials. Although the sur-
the fact that the presence of the hydrotalcite phase is necessary in face area (121 m2 g−1 ) of this material is relatively high and
86 A. R. Auxilio et al.

Table 1. Textural properties of Mg-Al-LDH-NO3 materials at different Mg:Al values

Mg:AlA Surface areaC Average pore Corrected pore volumeD Average densityE
[m2 g−1 ] diameterD [Å] [cm3 g−1 ] [g cm−3 ]

2.2 1.5 149 0.0058 2.00


3.1 1.4 152 0.0054 1.99
4.0 51 154 0.20 1.98
4.8 121 170 0.52 1.96
5.5 130 141 0.56 1.96
6.0 115 146 0.42 1.98
6.7 112 142 0.40 1.99
7.5 109 127 0.35 2.02
8.1 98 134 0.33 2.04
Mg(OH)2 B 14 134 0.045 2.33

A Mg:Al molar ratio determined by inductively coupled plasma–atomic emission spectroscopy analysis of the
layered double hydroxide (LDH).
B Pure magnesium hydroxide (included for comparison).
C Brunauer-Emmett-Teller method.
D Barrett-Joyner-Halenda method.
E Helium density.

it generally exhibits good affinity for the three anionic dyes, All isotherms for all anionic dyes adsorbing to Mg-Al-LH-
the surface looks flattened and the individual particles appear NO3 with a Mg:Al molar ratio between 3.1 and 6.0 (Fig. 6,
welded. The reasons for this are unclear at present. However, Accessory Publication Figs S2–S5) broadly exhibited a similar
with respect to adsorptive properties, no systematic correlation shape, i.e. they can be either classified as L- or H-type. There
can be drawn from this observation. This is partly because the appears to be a general trend as to the type of shape of the
adsorptive behaviour of these materials is different depending isotherm exhibited in going from a molar ratio of 2.2 to 6.0,
on what kind of anionic dye is used. viz. from S-type (moderate interaction; see Fig. 5a and c) to
an L-type (strong interaction; see Fig. S2a) and then H-type
(very strong interaction; e.g. Fig. 6a–c) and finally again an
Dye Adsorption Studies L-type (Figs S3–S5), although the traces in Figs S3–S5 are better
To study adsorption behaviour, adsorption isotherms were mea- described as L/H hybrid-type curves. In other words, adsorbate–
sured for the Mg-Al-LDH-NO3 material with a Mg:Al molar adsorbent chemisorption interactions seem to be at the optimum
ratio between 2.2 and 6.0, as these materials exhibited the best for the material with Mg:Al = 4.0. For the higher AY23 concen-
adsorption of ink-jet inks (see below). These studies applied trations, there is a general upward swing in the amount of AY23
commercially available anionic dyes of known composition and adsorbed on all the materials studied. Referring to the standard
structure (see Fig. 1). system of classification[33] of adsorption isotherms, this particu-
The equilibrium results for adsorption of the three dyes lar behaviour is observed over the first part of an S-type isotherm,
onto Mg-Al-LDH-NO3 at different divalent to trivalent metal suggesting some adsorbate–adsorbate interactions on the adsor-
ratios are shown in Figs 5, 6 and Accessory Publication bent that may be responsible for the so called ‘co-operative’
Figs S3–S7. According to their shape, one can obtain some adsorption mechanism.
insights into the processes and mechanisms taking place between The dye capacities (Cm ) and binding constants (Ka ) obtained
the liquid and solid interface. Of all the materials tested, from fitting of the data to the Langmuir equation are listed in
the adsorption isotherms (Fig. 5) of Mg-Al-LDH-NO3 with Table 2 for those systems where fitting was satisfactory. The
Mg:Al = 2.2 showed a peculiar behaviour towards the three dyes data reveal that Mg-Al-LDH-NO3 adsorbents exhibit different
used, in the sense that only one dye exhibited an isotherm that affinities for the three dyes at different Mg:Al molar ratios. The
could be fitted with the Langmuir model (AY23, Fig. 5b), but maximum Ka value of 1.1 × 106 M−1 was observed for the mate-
even in this case the fit was poor. Nevertheless, this material rial with Mg:Al = 4.0 for AB9. For AY23, Mg-Al-LDH-NO3 had
exhibited an H-type curve[33] expected of high-affinity mate- much the same affinity for this dye provided the Mg:Al molar
rials. The initial slope of the isotherm was almost vertical, ratio was between 2.2 and 4.0 (average Ka is 2.4 × 105 M−1 ).
suggesting that AY23 has a very high adsorption affinity towards Note, however, that for this dye, the adsorption behaviour is
this particular type of material. Following this sharp initial poorly fitted by the Langmuir equation (e.g. see Figs 5b and 6b).
rise, adsorption continued to increase but appeared to reach Caution is therefore required in the interpretation of the fitted
a plateau at higher dye concentrations, corresponding to satu- parameter. It is tempting to conclude that the optimum affinity
ration of available sites. However, AB9 and AR37 manifested of this type of material for the three anionic dyes is obtained with
another class of isotherm (Fig. 5a and c, respectively) that is the material having an Mg:Al ratio of 4.0. Note that the dye affin-
best described as an S-type according to the classification of ity test that will be discussed later revealed that BCI-Cyan ink
Giles et al.[33] This type of isotherm could not be fitted with has an excellent affinity for Mg-Al-LDH-NO3 with an Mg:Al
the Langmuir equation.[34] An S-type isotherm applied to this molar ratio between 4.8 and 5.5, BCI-Yellow ink has excellent
specific system means that there is a moderate intermolecular affinity for Mg-Al-LDH-NO3 with Mg:Al molar ratios between
attraction between the adsorbate and the adsorbent and that there 2.2 and 4.0, whereas BCI-Magenta (red) ink exhibits excel-
is strong competition for substrate sites, from molecules of the lent affinities for Mg-Al-LDH-NO3 materials with Mg:Al molar
solvent or of another adsorbed species. ratios between 4.8 and 5.5. The slight disagreement between
Adsorption of Ink-Jet Inks and Anionic Dyes 87

(a) (e)

500 nm 500 nm
Ratio  2.2 Ratio  5.5

(b) (f)

500 nm 500 nm
Ratio  3.1 Ratio  6.0

(c) (g)

500 nm 500 nm
Ratio  4.0 Ratio  6.7

(d) (h)

500 nm 500 nm
Ratio  4.8 Ratio  7.5

(i)

500 nm
Ratio  8.1

Fig. 4. Scanning electron micrographs of Mg-Al-LDH-NO3 with Mg:Al molar ratios of: (a) 2.2; (b) 3.1; (c) 4.0; (d) 4.8;
(e) 5.5; (f) 6.0; (g) 6.7; (h) 7.5; and (i) 8.1.

the adsorption isotherm studies using commercially available The effect of electrostatic interaction between the surface of
dyes (AB9, AY23, and AR37) and the dye affinity test using Mg-Al-LDH-NO3 and the dye is difficult to rationalize for this
commercially available ink-jet inks (BCI-Cyan, BCI-Yellow, and particular system.Theoretically,AY23 should give the highest Ka
BCI-Magenta) could be due to the effect of other additives (e.g. value, followed by AB9 and then AR37. This is because AY23
solvent, surfactant, biocide, buffer, chelating agent) present in and AB9 have three anionic groups, whereas AR37 has only
ink-jet inks. two. Furthermore, comparing AY23 and AB9, one of the anionic
88 A. R. Auxilio et al.

groups in AY23 is a carboxylate group, which is more basic than the surface of the adsorbent; (2) the morphology of the surface
sulfonates and would be expected to have higher affinity for Mg- of the adsorbent; and (3) multipoint interactions.
Al-LDH-NO3 . None of the adsorbent material studied exhibited In regard to the maximum monolayer coverage (Cm ), the data
this kind of trend. AB9 almost always gives the highest Ka value, reveal that the Mg-Al-LDH-NO3 adsorbent gives different Cm
followed by AR37 and then AY23. This observation suggests that values depending on the Mg:Al molar ratio and the adsorbate
dye affinity is not solely related to the electrostatic interaction used. For AB9, the highest Cm value (6.3 × 10−2 mmol g−1 )
but to other factors such as: (1) the stereochemistry of the dye on can be attained at Mg:Al = 5.5, whereas it is at Mg:Al = 3.1

(a) (a)
4.0
Experimental data
(not fitted) 7
3.5
Cs [mmol g1]  102

Cs [mmol g1]  102


6
3.0
5
2.5
4
2.0
3
1.5
Experimental data
2
Langmuir
1.0

0 2 4 6 8 0 2 4 6 8
Cp [mmol mL1]  104 Cp [mmol mL1]  104

(b) (b)

Experimental data
14 Langmuir 14
Cs [mmol g1]  102

Cs [mmol g1]  102

12
12

10
10
8
8
6 Experimental data
6 Langmuir

0 5 10 15 20 25 30 35 0 10 20 30 40
Cp [mmol mL1]  104 Cp [mmol mL1]  104

(c) (c)

Experimental data 9.0


(not fitted)
8.5
Cs [mmol g1]  102

3.5
Cs [mmol g1]  102

8.0

3.0 7.5

7.0
2.5
6.5
Experimental data
6.0 Langmuir
2.0 0

15 20 25 30 35 40 0 10 20 30
Cp [mmol mL1]  104 Cp [mmol mL1]  104

Fig. 5. Adsorption isotherm for (a) Acid Blue 9; (b) Acid Yellow 23; and Fig. 6. Adsorption isotherm for (a) Acid Blue 9; (b) Acid Yellow 23;
(c) Acid Red 37 binding to Mg-Al-LDH-NO3 with Mg:Al = 2.2. For (b), and (c) Acid Red 37 binding to Mg-Al-LDH-NO3 with Mg:Al = 4.0. Full
full line represents fitting of the data to the Langmuir equation. Cs , the lines represent fitting of the data to the Langmuir equation. Cs , the amount
amount of dye adsorbed to the surface; Cp is the amount of dye remaining of dye adsorbed to the surface; Cp is the amount of dye remaining in the
in the solution. solution.
Adsorption of Ink-Jet Inks and Anionic Dyes 89

and 4.0 for AY23 (Cm = 15 × 10−2 mmol g−1 ) and AR37 and AR37 adsorbates, which exhibited optimum adsorption
(Cm = 8.3 × 10−2 mmol g−1 ) respectively. For all the materials (Cm = 15.3 × 10−2 mmol g−1 ) forAY23 at Mg:Al = 3.1 (surface
studied, the general trend for the Cm value is that for all the mate- area (SA) = 1.4 m2 g−1 ) and at Mg:Al = 4.0 (SA = 51 m2 g−1 )
rials studied, AY23 gave the highest value, followed by AR37 and for AR37 (Cm = 8.3 × 10−2 mmol g−1 ). Two explanations can
then AB9. be proposed here: (1) different kinds of interaction can take
Previously,[14] it was concluded that the adsorptive capacity place between the adsorbate and the adsorbent owing to the
(using AB9) of Mg-Al-LDH-CO3 material is a direct conse- different number and type of anionic groups present in the
quence of surface area. Thus, the Mg-Al-LDH-CO3 material dye; and (2) even at the relatively low concentrations (5 mM)
with Mg:Al = 3.1 had the highest surface area (123 m2 g−1 ) and for AY23 (3− charge) and AR37 (2− charge) used in this
the highest Cm value (7.6 × 10−2 mmol g−1 ). This also appears study, these dyes have the potential to replace the nitrates (1−
to hold true for the Mg-Al-LDH-NO3 material provided the same charge) in the interlamellar region (intercalate) mainly because
dye (AB9) is used (see Adsorption Isotherm section). In fact, of their higher charge. However, this requires further study.
for AB9, optimum adsorption (Cm = 6.3 × 10−2 mmol g−1 ) was It is possible that more dye is adsorbed by materials with
observed at Mg:Al = 5.5, where the surface area (130 m2 g−1 ) higher layer charge density (lower Mg:Al).[35] Note that the
is at its maximum. With respect to pore size, no correlation with work of Ulibarri and Hermosin and their coworkers[16] seems
adsorptive capacity can be made with this property. As the pore in contradiction with the present study as discussed in our pre-
volume changes in accordance with the surface area, the same vious publication.[14] Changing the Mg:Al molar ratio clearly
conclusion derived from surface area with respect to the adsorp- affects the adsorptive capacity of Mg-Al-CO3 material. But,
tive capacity can be expected. Furthermore, the average density the fact that Ulibarri and Hermosin reported that the adsorp-
does not influence the adsorptive behaviour of Mg-Al-LDH- tion behaviour of 2,4,6-trinitrophenol on the Mg-Al-Cl system
NO3 materials as it does not change significantly as the Mg:Al was affected by the variation of the Mg:Al molar ratio seems
molar ratio changes. The average density (2 g cm−3 ), however, to suggest the likelihood of intercalation of AY23 and AR37
for Mg-Al-LDH-NO3 materials is the same as for the Mg-Al- in our study, as mentioned above. However, the work of Fetter
LDH-CO3 materials reported previously.[14] Therefore, it can et al.[17] is in agreement with findings presented herein. This
also be concluded that changing the interlayer anions from CO2− 3 shows that adsorption behavioural studies using hydrotalcite-
to NO− 3 does not affect the density of the material.
like materials are complex and can involve many factors such
It should be emphasized here that above and below a molar as: (1) the procedure of sorption studies; (2) the mechanism
ratio of 5.5 for Mg-Al-LDH-NO3 , the surface area decreases of adsorption (surface or interlamellar); and (3) the type of
as does the adsorptive capacity for AB9. This applies to AY23 sorbate.

Table 2. Langmuir model parameters for Acid Blue 9 (AB9), Acid Yellow 23 (AY23), and Acid Red 37 (AR37) binding to Mg-Al-LDH-NO3 system
at ratios 2.2, 3.1, 4.0, 4.8, 5.5, and 6.0
Bold entries represent the highest measured Cm and Ka values

Mg:Al molar ratio Cm [mmol g−1 ] Ka [M−1 ]


AB9 AY23 AR37 AB9 AY23 AY37

2.2 – 12.0 (±0.4) × 10−2 – – 2.53 (±0.72) × 105 –


3.1 – 15.3 (±1.2) × 10−2 – – 2.76 (±0.84) × 105 –
4.0 5.53 (±0.33) × 10−2 11.9 (±0.65) × 10−2 8.29 (±0.20) × 10−2 1.10 (±0.35) × 106 2.02 (±0.75) × 105 1.44 (±0.45) × 105
4.8 5.59 (±0.32) × 10−2 – 8.22 (±0.17) × 10−2 2.68 (±0.80) × 105 – 5.3 (±1.4) × 104
5.5 6.30 (±0.29) × 10−2 – 10.2 (±0.37) × 10−2 1.75 (±0.37) × 105 – 6.5 (±2.4) × 104
6.0 4.40 (±0.32) × 10−2 – 6.40 (±0.20) × 10−2 5.4 (±2.0) × 104 – 10.0 (±3.4) × 104

100
90
80
Dye adsorbed [%]

70
60
50
40
30 BCI-Magenta
20 BCI-Yellow
10 BCI-Cyan
0
2.2 3.1 4.0 4.8 5.5 6.0 6.7 7.5 8.1
Mg:Al molar ratio

Fig. 7. Percentage of commercial ink-jet inks (BCI-Cyan, BCI-Yellow, and BCI-Magenta) adsorbed onto the surface of Mg-Al-LDH-
NO3 at different Mg:Al molar ratios. Note: adsorbent mass = 0.5 g, T = ambient, pH of 10% ink solutions (20% v/v glycerol/water
solvent) are as follows: BCI-Cyan = 8.1, BCI-Yellow = 4.7, and BCI-Magenta = 5.4.
90 A. R. Auxilio et al.

A qualitative dye affinity test was also conducted for the com- at pH 9.50 (±0.05) by the modification of previously described
mercial ink-jet inks to initially investigate the viability of these methods.[14,26] A typical synthetic procedure leading to a sam-
materials to be used in ink-receptive layer (IRL) formulations. ple with Mg:Al = 2.2 (actual Mg:Al molar ratio added is 2.0)
The percentages of dye adsorbed for each Mg-Al-LDH-NO3 involved the following.
sample with different Mg:Al ratios are shown in Fig. 7. At Solution A was prepared by dissolving Mg(NO3 )2 ·6H2 O
the outset, the materials tested have shown some affinity for (61.54 g, Sigma–Aldrich) andAl(NO3 )3 ·9H2 O (45.02 g, Sigma–
the dye in solution of varying degree, spanning from 19% for Aldrich) in water (150 mL). Solution B was prepared by dis-
BCI-Magenta with Mg-Al-LDH-NO3 of Mg:Al = 3.1 to 96% solving NaOH (23.27 g, Merck) in water (150 mL). Solutions
for BCI-Yellow with Mg-Al-LDH-NO3 of Mg:Al = 2.2. BCI- A and B were added dropwise to a four-necked round-bottom
Yellow exhibited a different adsorption behaviour with respect flask containing freshly boiled water (120 mL) under a nitrogen
to Mg:Al molar ratios of Mg-Al-LDH-NO3 compared with atmosphere. The addition of solutions A and B was performed
BCI-Cyan and BCI-Magenta. with constant stirring (using a magnetic stirrer) and at ambi-
The highest affinity for BCI-Yellow was found for Mg-Al- ent temperature, and at such a rate that the pH of the reaction
LDH-NO3 materials with an Mg:Al molar ratio between 2.2 mixture was maintained at 9.5 (±0.05) using a Metrohm 736 GP
and 4.0, whereas for BCI-Cyan and BCI-Magenta, a high affin- Titrino autotitrator. The resulting slurry was aged at 100◦ C using
ity could be achieved with an Mg:Al molar ratio between 4.8 an oil bath for 16 h under a nitrogen atmosphere. The precipi-
and 5.5. Within the Mg:Al molar ratios studied, these materials tate was filtered (no. 3 sintered filter funnel), washed twice with
were found to adsorb slightly more yellow dye than magenta deionized and decarbonated water and dried at 100◦ C for 24 h.
and cyan. This may be due to the yellow dye having more acidic All products were then ground with an agate mortar and pestle
(anionic) groups available to interact with the cationic surface. before characterization.
Mg-Al LDHs are well known to have a very high isoelectric point
(between 11 and 12).[36–38] Therefore, the surface is positively Preparation of Ink-Jet Inks and Anionic Dye Solutions
charged at the pH of ink samples. Note that the pH of solu- Commercial inks (BCI-Cyan, BCI-Yellow, and BCI-Magenta)
tions of BCI-Yellow, BCI-Magenta, and BCI-Cyan are 4.7, 5.4, were diluted using 20% (v/v) glycerol/deionized water, giving a
and 8.1, respectively. Currently, studies are being conducted by 10% (v/v) ink solution before the dye affinity test. Glycerol is
the authors aimed at understanding the fundamental adsorption one of the common solvents used in the formulation of ink-jet
mechanisms by taking into account the dye structures. inks.[9] These inks were sold as refillers for Canon ink-jet printer
Conclusion cartridges and are used in their mid-range ink-jet printers.
Commercially available dyes (Fig. 1), namely CI AB9 (AB9,
As part of ongoing efforts aimed at developing inorganic pig-
Saujanya Dyechem), CI Y23 (AY23, Sigma–Aldrich), and CI
ments for application as media for ink-jet printing, we have
AR37 (AR37, Sigma–Aldrich), were used to prepare dye solu-
investigated the dye-adsorption properties of a series of Mg-Al-
tions consisting of 1 mM AB9, 5 mM AY23, and 5 mM AR37
LDH-NO3 materials with different Mg:Al molar ratios. Equi-
using pure water (Milli-Q) taking into account the dye content
librium adsorption experiments, using three anionic dyes, AB9,
from the as-received dyes. These dye concentrations were based
AY23, and AR37, indicated a very strong interaction between the
on commercial ink-jet ink concentrations mentioned above (note
adsorbate and adsorbent for several Mg:Al ratios. The highest
that AB9 has a higher ε value). Table 3 lists the wavelength
binding affinities (Ka ), were observed for AB9 and AY23 and for
of maximum absorbance (λmax ) and molar absorptivity (ε) for
Mg:Al = 4.0. We note that the Langmuir equation fitted many of
each dye.
the adsorption isotherms poorly, indicating that multiple adsorp-
tion mechanisms may be in operation for these materials. This
Characterization Techniques
is in contrast with the behaviour of the previously reported Mg-
Al-LDH-CO3 system. In general, surface adsorption processes Powder X-ray diffraction analyses were performed using a
showed no systematic correlation with the textural properties Philips 1140 diffractometer under the following conditions:
of the materials. However, it can be generalized that large pore 40 kV, 25 mA, CuKα (λ = 0.15406 nm) radiation. The samples,
sizes and a pure LDH phase are necessary prerequisites for this as unoriented powders, were scanned in steps of 0.02◦ (2θ) in
material to behave as an excellent adsorbent for the dyes used in the range from 2◦ to 70◦ at a speed of 2◦ per min.
this study. Metal analyses were performed via inductively coupled
Comparison of the adsorptive behaviour between Mg-Al- plasma–atomic emission spectroscopy by ALS Environmental,
LDH-NO3 with different Mg:Al molar ratios and ink-jet inks Victoria, Australia.
(BCI-Cyan, BCI-Yellow, and BCI-Magenta) as adsorbate indi- Infrared experiments were carried out using an Equinox–IFS
cated the ability of these dyes to adsorb varies with the material 55 spectrometer. Samples were mixed with KBr to produce a
composition. BCI-Yellow was found to have high affinity for disc with a 1% compound loading. The spectra were recorded
Mg-Al-LDH-NO3 with an Mg:Al molar ratio between 2.0 and between 4000 and 400 cm−1 at a resolution of 2 cm−1 .
4.0, while BCI-Cyan and BCI-Magenta showed a high affinity
when the Mg:Al molar ratio was between 4.8 and 5.5. Physi- Table 3. Some properties of commercial dyes used in this work
cal mixing of appropriate proportions of these materials could,
therefore, be used to produce a composite material with excellent Name Dye content Molecular mass εexp A λmax A
affinity for all commercial ink-jet inks. [%] [g mol−1 ] [M cm−1 ]
−1 [nm]

Experimental AB9 85 792.85 1.37 × 105 629


AY23 70 534.36 2.57 × 104 426
Preparation of Adsorbent Material
AR37 90 514.53 1.88 × 104 513
Mg-Al LDHs with nitrates as interlayer ions (Mg-Al-LDH-
NO3 ) with varying Mg2+ :Al3+ ratios (2.2–8.1) were prepared A From UV-visible spectral analysis.[24]
Adsorption of Ink-Jet Inks and Anionic Dyes 91

Surface morphologies and particle size were studied using [11] D. Neumann, S. Loffler, W. Raverty, N. Vanderhoek, P. Andrews,
a Philips XL30 field emission scanning electron microscope P. Junk, L. Spiccia, Pigments for Ink Receptor Layers – Characteri-
manufactured by Oxford Instruments operating at 5 kV. sation and Assessment. 58th Appita Conference Proceedings 2004, p.
Surface area, pore size, and pore volume were studied using 383 (Appita: Melbourne).
[12] E. Svanholm, Printability and ink-coating interactions in inkjet
an ASAP 2010 Micromeritics instrument. The samples were
printing, Ph.D. Thesis, Karlstad University, Sweden, 2007.
degassed at 378 K before analyzing using nitrogen adsorption–
[13] N. Iyi, T. Matsumoto, Y. Kaneko, K. Kitamura, Chem. Mater. 2004,
desorption at 77 K. 16, 2926. doi:10.1021/CM049579G
The density was determined by helium pycnometry using an [14] A. R. Auxilio, P. C. Andrews, P. C. Junk, L. Spiccia,
AccuPyc 1330 gas-displacement pycnometer manufactured by D. Neumann, W. Raverty, N. Vanderhoek, Polyhedron 2007, 26, 3479.
Micromeritics. doi:10.1016/J.POLY.2007.03.019
[15] A. D. Roy, C. Forano, J. P. Besse, Layered Double Hydroxides: Synthe-
Adsorption Experiments sis and Post-Synthesis Modification, in Layered Double Hydroxides:
Present and Future (Ed. V. Rives) 2001, pp. 1–37 (Nova Science
Mg-Al-LDH-NO3 materials were subjected to a dye affin-
Publishers: New York, NY).
ity test using purchased ink-jet inks according to published [16] M. A. Ulibarri, M. D. C. Hermosin, Layered Double Hydroxides in
methodology.[14] Ink-jet ink solutions (5 mL, 10% v/v in 20% Water Decontamination, in Layered Double Hydroxides: Present and
v/v glycerol/water solvent) were added to the pigment (0.5 g) in Future (Ed. V. Rives) 2001, pp. 251–284 (Nova Science Publishers:
a centrifuge tube. The tube was agitated using a vortex mixer New York, NY).
(2500 rpm) for 1 min followed by centrifugation (3000 rpm, [17] G. Fetter, M.T. Olguin, P. Bosch,V. H. Lara, S. Bulbulian, J. Radioanal.
10 min, ambient temperature). The supernatant liquid was care- Nucl. Chem. 1999, 241, 595. doi:10.1007/BF02347218
fully removed and an aliquot (25 µL) was diluted with distilled [18] K. Zou, H. Zhang, X. Duan, Chem. Eng. Sci. 2007, 62, 2022.
water (5 mL) and analyzed using a Varian Cary 300 Bio UV- doi:10.1016/J.CES.2006.12.041
[19] Y. Wang, H. Gao, J. Col. Inter. Sci. 2006, 301, 19. doi:10.1016/J.JCIS.
visible spectrophotometer. For materials that exhibited very high
2006.04.061
affinity for the dye, a different dilution factor was used so that
[20] Y.-F. Chao, P.-C. Chen, S.-L. Wang, Appl. Clay Sci. 2008, 40, 193.
the absorbance reading was in a measurable range. doi:10.1016/J.CLAY.2007.09.003
For Mg-Al-LDH-NO3 materials (Mg:Al = 2.2 to 6.0) with [21] S. Vreysen, A. Maes, Appl. Clay Sci. 2008, 38, 237. doi:10.1016/
the best adsorption of ink-jet inks, adsorption isotherms were J.CLAY.2007.02.010
determined for commercial dyes of known composition and [22] M. del Arco, S. Gutierrez, C. Martin, V. Rives, J. Rocha, J. Solid State
structure (approximate pH of measurement 7–8) as discussed Chem. 2000, 151, 272. doi:10.1006/JSSC.2000.8653
in previous publications.[14,23,24] Essentially, this involved using [23] A. R. Auxilio, P. C. Andrews, P. C. Junk, L. Spiccia, Dyes Pigments
a similar procedure as outlined in the adsorption experiments, 2009, 81, 103. doi:10.1016/J.DYEPIG.2008.09.011
except that varying amounts of adsorbent materials were exposed [24] A. R. Auxilio, P. C. Andrews, P. C. Junk, L. Spiccia, D. Neumann,
W. Raverty, N. Vanderhoek, J. M. Pringle, J. Mater. Chem. 2008, 18,
to a constant dye concentration.
2466. doi:10.1039/B715545J
[25] Y.-F. Chao, P.-C. Chen, S.-L. Wang, Appl. Clay Sci. 2008, 40, 193.
Accessory Publication doi:10.1016/J.CLAY.2007.09.003
The Accessory Publication contains basal spacings derived [26] S. Miyata, Clays Clay Miner. 1975, 23, 369. doi:10.1346/CCMN.
from the XRD patterns (Table S1), N2 adsorption–desorption 1975.0230508
isotherms (Fig. S1) and dye adsorption isotherms (Figs S2– [27] S. Miyata, Clays Clay Miner. 1983, 31, 305. doi:10.1346/CCMN.
1983.0310409
S5) for Mg-Al-LDH-NO3 with different Mg:Al ratios. The
[28] S.-L. Wang, P.-C. Wang, Coll. Surf. A: Physicochem. Eng. Aspects
Accessory Publication is on the Journal’s website.
2007, 292, 131. doi:10.1016/J.COLSURFA.2006.06.014
[29] A. Vaccari, Appl. Clay Sci. 1999, 14, 161. doi:10.1016/S0169-
Acknowledgements 1317(98)00058-1
The authors wish to thank the Smartprint Cooperative Research Centre, [30] Z. P. Xu, H. C. Zeng, J. Phys. Chem. B 2001, 105, 1743. doi:10.1021/
Australia, for the financial support during the course of this study. A.R.A. JP0029257
is also grateful to Monash Research Graduate School for a Postgraduate [31] R. Marangoni, L. P. Ramos, F. Wypych, J. Col. Inter. Sci. 2009, 330,
Publication Award. 303. doi:10.1016/J.JCIS.2008.10.081
[32] K. S. W. Sing, D. H. Everett, R. A. W. Haul, L. Moscou, R. A. Pierotti,
References J. Rouquerol, T. Siemieniewska, Pure Appl. Chem. 1985, 57, 603.
doi:10.1351/PAC198557040603
[1] Y. Arai, US Patent 20070176991 2007.
[33] C. H. Giles, T. H. MacEwan, S. N. Nakhwa, D. Smith, J. Chem. Soc.
[2] L. M. Carreira, K.-J. Brodsky, M. A. Evans, A. L. Wickett, C. C. Lyons,
1960, 3973. doi:10.1039/JR9600003973
US Patent 5925177 1999.
[34] V. P. Evangelou, Environmental Soil and Water Chemistry 1998 (John
[3] W. Fassler, C. D. DeBoer, J. E. Mooney, US Patent 6059869 2000.
Wiley & Sons: New York, NY).
[4] K. Ito, Y. Fujiwara, US Patent 5560996 1996.
[35] Y. F. Chao, J. J. Lee, S. L. Wang, J. Hazard. Mater. 2009, 165, 846.
[5] K. Ito, Y. Fujiwara, US Patent 6281270 B1 2001.
doi:10.1016/J.JHAZMAT.2008.10.063
[6] C. Jackson, K. H. Kung, R. D. Bauer, US Patent 20050235867 2005.
[36] R. R. Delgado, M. A. Vidaurre, C. P. D. Pauli, M. A. Ulibarri,
[7] F. Shi, P. Doll, W. Wnek, M. Andreottola, US Patent 6641257 2003.
M. J. Avena, J. Coll. Inter. Sci. 2004, 431.
[8] J. E. Valentini, G. P. Morris, S. L. Issler, J. W. Wheeler, US Patent
[37] O. C. Wilson, T. Olorunyolemi, A. Jaworski, L. Borum, D. Young,
2007/0058016 AI 2007.
E. Dickens, C. Oriakhi, Cer. Trans. 2000, 110, 93.
[9] P. Gregory, High-Technology Applications of Organic Colorants 1991
[38] J. O. C. Wilson, T. Olorunyolemi, A. Jaworski, L. Borum, D. Young,
(Plenum Press: New York, NY).
A. Siriwat, E. Dickens, C. Oriakhi, M. Lerner, Appl. Clay Sci. 1999,
[10] M. I. El-Barghouthi, A. H. El-Sheikh, Y. S. Al-Degs, G. M.
15, 265. doi:10.1016/S0169-1317(99)00023-X
Walker, Separation Sci. Tech. 2007, 42, 2195. doi:10.1080/
01496390701444030

You might also like