You are on page 1of 9

Journal of Colloid and Interface Science 406 (2013) 51–59

Contents lists available at SciVerse ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Mechanism and kinetics of nanostructure evolution during early stages


of resorcinol–formaldehyde polymerisation
Katarzyna Z. Gaca ⇑, Jan Sefcik ⇑
Department of Chemical and Process Engineering, University of Strathclyde, 75 Montrose Street, G1 1XJ Glasgow, UK

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

Article history: Resorcinol and formaldehyde react in aqueous solutions to form nanoporous organic gels well suited for a
Received 17 February 2013 wide range of applications from supercapacitors and batteries to adsorbents and catalyst supports. In this
Accepted 2 May 2013 work, we investigated the mechanism and kinetics of formation of primary clusters in the early stages of
Available online 3 June 2013
formation of resorcinol–formaldehyde gels in the presence of dissolved sodium carbonate. Dynamic Light
Scattering measurements showed that size of freely diffusing primary clusters was independent of both
Keywords: reactant and carbonate concentrations at a given temperature, reaching the mean hydrodynamic radius
Sol–gel
of several nanometres before further changes were observed. However, more primary clusters formed at
Gelation
Resorcinol–formaldehyde
higher carbonate concentrations, and cluster numbers were steadily increasing over time. Our results
Organic gels indicate that the size of primary clusters appears to be thermodynamically controlled, where a solubil-
Dynamic Light Scattering ity/miscibility limit is reached due to formation of certain reaction intermediates resulting in approxi-
Primary clusters mately monodisperse primary clusters, most likely liquid-like, similar to formation of micelles or
Spontaneous emulsification spontaneous nanoemulsions. Primary clusters eventually form a particulate network through subsequent
Nanoemulsions aggregation and/or coalescence and further polymerisation, leading to nanoscale morphologies of result-
Thermodynamic control ing wet gels. Analogous formation mechanisms have been previously proposed for several polymerisation
and sol–gel systems, including monodisperse silica, organosilicates and zeolites.
Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction been widely used in materials synthesis [8], and although it has
been traditionally associated with the synthesis of inorganic mate-
Organic gels based on formaldehyde and resorcinol were first rials, a wide range of organically modified or purely organic mate-
synthesised in the late 1980s [1] through an organic sol–gel pro- rials have been produced in this way [2]. The organic sol–gel
cess, and since then, they have become increasingly important in process is a convenient and environmentally friendly synthesis
a wide range of applications [1–3]. Due to their electrical conduc- method, where dissolved monomers undergo series of reactions,
tivity, very large surface areas and internal porosity, resorcinol– forming primary particles or clusters at nanometre to micrometre
formaldehyde aerogels have high potential as superior materials scale, which then subsequently coalesce and/or aggregate forming
for electrodes in electrochemical double-layer supercapacitors a system-spanning network immersed in the solvent: a wet gel.
and batteries [4,5], for capacitive deionisation units, as well as The resulting organic gel is then dried, carbonised and chemically
materials for hydrogen storage [6]. Therefore, further development activated. Additional treatments like solvent exchange and ageing
and tailoring of these materials for specific applications is of great may also be introduced prior to drying, in order to achieve a gel
importance for energy storage and low carbon technologies. Fur- with properties appropriate for a given application. Although there
ther important applications of carbon aerogels include catalyst exists considerable empirical experience regarding the synthesis of
supports and adsorbents [7]. organic gels, fundamental understanding of their nanostructure
Previous research has shown that resorcinol–formaldehyde gel evolution is lacking, and therefore, our ability to rationally tailor
properties can be tailored to specific requirements, such as pore their structure for specific applications has been limited.
surface area and pore size distribution, by changing the conditions The widely accepted reaction mechanisms for the synthesis of
of their synthesis in the sol–gel process [2]. The sol–gel process has resorcinol–formaldehyde polymers are shown in Fig. 1. It can be
divided into two steps: substitution of resorcinol with formalde-
hyde to form (simply, doubly or triply) substituted hydroxymethyl
⇑ Corresponding authors. Present address: Faculty of Chemical Engineering and resorcinol and step-growth polymerisation of substituted resor-
Technology, Cracow University of Technology, ul. Warszawska 24, 31-155 Cracow,
cinol [1,3].
Poland (K.Z. Gaca).
E-mail addresses: katarzyna@gaca.cat (K.Z. Gaca), jan.sefcik@strath.ac.uk (J.
Although resorcinol–formaldehyde polymerisation can take
Sefcik). place in aqueous solutions at room temperature without any added

0021-9797/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2013.05.062
52 K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59

Fig. 1. Overall reaction mechanism of resorcinol–formaldehyde polymerisation.

catalyst, the overall rate is relatively low [9], and it can be polymerising resorcinol–formaldehyde systems: microphase sepa-
expected, by analogy with a well-known phenol–formaldehyde ration [13,14] or aggregation of primary colloidal particles [15].
system, that the substitution reaction is both acid or base catalysed The microphase separation scenario is based on the idea of poly-
[1,3,10]. The polymerisation step, where the hydroxymethyl group meric chains growing too large and becoming insoluble and there-
reacts with another hydroxymethyl group, forming an ether bridge fore triggering demixing, resulting in two interpenetrating phases,
–CH2–O–CH2–, or with an unsubstituted resorcinol site, forming a one polymer rich and the other solvent rich, a process well known
methylene bridge –CH2–, can also be expected to be both acid and from other polymer systems. The colloidal aggregation scenario is
base catalysed [10]. Polymerisations are typically performed at ele- based on the idea of formation of primary colloidal particles which
vated temperatures (up to 90 °C) in order to speed up the process. then subsequently aggregate and form a space-filling network of
Moreover, previous research showed that in order to obtain very interconnected clusters. Recent analysis of Small Angle X-ray Scat-
high pore surface areas of resulting gels, neutral or slightly basic tering (SAXS) measurements from resorcinol–formaldehyde gels
conditions within a relatively narrow range of pH values are re- suggests that these two scenarios are unlikely to be distinguishable
quired [10], and this is often achieved by adding weak bases such from the resulting gel structures and in any case may well be two
as sodium carbonate as reaction catalysts to resorcinol–formalde- extremes of a more general process [16]. From published transmis-
hyde solutions. More recent research indicated that not all bases sion electron micrographs [13], it is apparent that the dry network
are equally effective at same pH values in promoting high surface of resorcinol–formaldehyde gels produced in the presence of car-
areas of resulting gels and that both concentration and type (anion bonates consists of interconnected dense primary particles of a
and cation) of basic catalysts added to modify solution pH affect few to a few tens on nanometres, with smaller primary particles
the properties of the final gel [11], suggesting that the role of these for higher carbonate concentrations. However, it is not clear what
catalysts is beyond just providing basic conditions. Electrolytes, is a mechanism of formation of these primary particles and how is
such as sodium carbonate, are well known to destabilise colloidal it controlled by solution pH and the nature of cations and anions
suspensions due to screening repulsive electrostatic interactions used. Formation of primary clusters was previously investigated
between colloidal particles. Also, electrolytes are known to induce by SAXS [17], where clusters with radius of gyration of 2–4 nm
liquid–liquid phase separation in water–alcohol mixtures leading were observed before the onset of gelation at 25 °C. In a recent
to colloidal emulsions. study, SAXS and X-ray photon correlation spectroscopy was used
We note that in aqueous solutions, formaldehyde is mainly to monitor gel formation at 70 °C [18], where it was observed that
present in its hydrated form as methylene glycol and its oligo- primary clusters were gradually growing with their radius of gyra-
mers/polymers, and commercial formaldehyde solutions are nor- tion from about 2 to about 10 nm, after which development of
mally stabilised with methanol in order to prevent formaldehyde increasingly stiff network dynamics was observed. Sol–gel transi-
polymerisation, so that methylene glycols are accompanied by tion process in resorcinol–formaldehyde solutions at 25 °C was
methoxymethylene glycol and their small oligomers, while the investigated by Dynamic Light Scattering, where monomodal de-
concentration of formaldehyde in its aldehyde form is in fact very cay time distributions were observed initially with apparent clus-
low (usually not exceeding 100 ppm) [12]. Therefore, the initial ter hydrodynamic diameters growing gradually up to about
reaction mixture in resorcinol–formaldehyde polymerisation is 5 nm. The initial formation of these primary clusters was followed
an aqueous solution of resorcinol and methylene glycols in the by the development of much slower autocorrelation function de-
presence of strong electrolytes (such as sodium carbonate) added cay modes, which were presented in term of steeply increasing
as catalysts. As reactions proceed, a range of products starting with apparent hydrodynamic diameters of growing colloidal particles
variably substituted resorcinol and their condensation products [15].
appear over time, with unknown solubility/miscibility properties The nanoscale structure of resorcinol–formaldehyde gels as ob-
in respect to the background solution at relevant temperatures. served by electron microscopy appears to be morphologically sim-
While it is expected that sufficiently long polymer chains will ilar to the one of silica-based gels from silicon alkoxides through
eventually become insoluble and undergo microphase separation, hydrolysis and subsequent step-growth polymerisation in near
it may well be possible that solubility/miscibility limits are neutral or slightly basic aqueous solutions, where the sol–gel pro-
reached earlier in the polymerisation process as relatively small cess is similar in some respects to synthesis of resorcinol–formal-
molecular weight intermediates and oligomers are produced. dehyde gels. It has been previously proposed that formation of
There are two principal theories which can be found in the lit- primary particles in silica sol–gel systems may include formation
erature aiming to explain the mechanism of the gel formation in of transient nanoemulsions through molecular demixing driven
K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59 53

by production of small molecular weight reaction intermediates was then added to the beaker with dissolved resorcinol, and after
before any substantial polymerisation commences [19]. Similar 10 min of further stirring to dissolve sodium carbonate, the
mechanisms based on formation of nanoscale liquid-like interme- required amount of formaldehyde solution (F solution) was added
diates have been proposed for several polymerisation and sol–gel to the mixture and stirring continued for another 30 min. The solu-
systems, including monodisperse Stoeber silica particles [20], tions were then filtered in two steps, first using 0.45 lm and then
zeolites [21] and organosilicates [22,23]. We note that while small 0.02 lm pore size syringe filter, directly into clean vials pre-
silica/organo-silica oligomers appear to be poorly miscible/soluble washed with deionised water and filtered resorcinol–formalde-
in water–alcohol mixtures at basic and neutral pH conditions, hyde solution. Vials were then sealed with a cap and parafilm,
resulting in formation of colloidal primary particles, there appear placed in the DLS measurement cell where they were thermostated
to be no solubility issues in acidic solutions where transparent gels at 55 °C and measurements were taken at certain time intervals.
are formed without any apparent nanoscale phase separation Since DLS measurements were not possible to do in situ at 80 °C
phenomena. with the instrument available, they were conducted in an alterna-
Interestingly, it appears that the mechanism of resorcinol– tive manner. Solution was divided upon filtration into several par-
formaldehyde gel formation under basic conditions is also different allel samples which were placed in an electric oven preheated to
from that in acidic conditions, where polymerisation-induced 80 °C. The vials were then taken out from the oven at certain time
critical opalescence, as inferred from SAXS measurements [24], intervals, rapidly cooled down to ambient temperature in an ice
has been proposed to be linked to polymerisation-induced phase bath to quench the reaction and taken for DLS measurements per-
separation, leading to formation of micron-sized particles formed at 25 °C.
subsequently aggregating and forming a gel. Nevertheless, the In Dynamic Light Scattering experiments, autocorrelation func-
mechanism of formation of primary clusters in resorcinol–formal- tions g1(s) were measured using a digital correlator (ALV/LSE-
dehyde polymerisation under basic and neutral conditions and the 5004). The wavelength of incident laser light was k = 632.8 nm,
role(s) played by basic catalysts, such as sodium carbonate, are still and the scattering angle h was 90°. When the autocorrelation decay
poorly understood. was showing an exponential decay, indicating unhindered Brown-
The objective of this work was to investigate mechanism and ian diffusion, we used the cumulant method [26] to estimate in the
kinetics of formation of primary clusters during early stages of res- initial decay rate C [s 1]. From this, one can determine the mean
orcinol–formaldehyde polymerisation in the presence of sodium diffusion coefficient D, using C = Dq2, where q = (4p/k)sin(h/2),
carbonate by analysing effects of carbonate and reactant concen- and the mean hydrodynamic radius Rh, using the Stokes–Einstein
trations and temperature, using Dynamic Light Scattering to mon- equation, D = kBT/6plRh, where kB is the Boltzmann constant, T is
itor formation of primary clusters and subsequent structure the absolute temperature, and l is dynamic viscosity (taken here
development towards gel formation. to be equal to that of pure water at a given temperature).
For estimation of gelation times, solutions were transferred into
25 ml PP flasks (diameter 25 mm, height 100 mm), sealed and
2. Experimental placed in an electric oven preheated to a desired temperature.
The process of gelation was monitored by visual observations of
Resorcinol (99 wt.%), formaldehyde (37 wt.%, aqueous solution changes in the solution flowability, and gelation time was taken
stabilised by 13 wt.% methanol; the concentration of methanol here as time measured from the moment a sample of reacting
was determined experimentally using the NMR spectroscopy mixture is placed in an electric oven set at the desired temperature
[25]) and sodium carbonate (P99.5 wt.%, anhydrous ACS reagent) and periodically checked until a lack of flow at tilting by 45° is ob-
were all purchased from Sigma–Aldrich. Syringe filters (0.45 lm served. Although this kind of gel time estimation is subject to sev-
pore size, PTFE membrane and PP housing; 0.02 lm pore size, Ano- eral factors, including the diameter of the vessel and the total
top 10; manufactured by Whatman) were purchased from Fisher volume of the solution, it is a suitable tool for the assessment of
Scientific. Filtration was carried out with rubber-free PVC/PP syrin- differences in gel formation kinetics among similar samples sub-
ges. Vials for DLS experiments were made from borosilicate glass, ject to different reaction conditions. It has been reported that there
with diameter of 10 mm and height 75 mm, and were also pur- has been a reasonable correlation between visual and rheological
chased from Fisher Scientific. Filtration was performed in order determination of gel times in similar resorcinol–formaldehyde sys-
to eliminate impurities, including dust, which could strongly affect tems, where visually estimated gel times were 5–45% longer than
DLS results. The reacting mixtures were filtered after the initial those determined rheologically [27]. With this caveat, we can use
mixing, but before thermal treatment, where only unsubstituted visual estimation of gel times for approximate determination of
and substituted resorcinol is present, but no larger oligomers are time windows for DLS measurement before networking of primary
formed under these conditions [25]. clusters and onset of gelation.
The compositions of resorcinol–formaldehyde solutions used
for sol–gel processes are customarily defined by ratios of amounts
of reagents, carbonate and water used. The most frequently used 3. Results and discussion
ratios are R/C (resorcinol to catalyst; mol/mol), R/F (resorcinol to
formaldehyde; mol/mol) and R/W (resorcinol to water; g/ml; W re- Initial gelation time observations were performed to estimate
fers to water added as a solvent, not the water present in formal- time scales appropriate for DLS measurements in order to monitor
dehyde solution). In all experiments, resorcinol to formaldehyde formation of primary clusters well before the gel formation itself,
ratio was kept constant at 0.5 mol/mol, while resorcinol to water since previous scattering studies typically focused on later stages
and overall carbonate concentrations were varied. The composi- of the process where extensive aggregation and/or cross-linking
tions of mixtures used in this work are shown in Table 1. Please of primary clusters resulted in complicated interplay of cluster
note that F in the R/F ratio refers to mols of formaldehyde, while size, structure and dynamics. Fig. 2 shows gelation times deter-
F solution refers to the volume of the added formaldehyde solution. mined for a range of carbonate concentrations at two different
All resorcinol–formaldehyde solutions were prepared as fol- temperatures. Here, ratios of resorcinol to formaldehyde (R/F)
lows. The required amounts of deionised water (W) and resorcinol and resorcinol to water (R/W) were kept constant at 0.5 mol/mol
(R) were transferred to a beaker and stirred using a magnetic stir- and 0.10 g/ml, respectively, while concentration of carbonate,
rer for 5 min until resorcinol fully dissolved. Sodium carbonate (C) expressed as the ratio of resorcinol to catalyst (R/C), was varied
54 K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59

Table 1
Compositions of investigated reaction mixtures.

R/W R/C (mol/mol) R/F (mol/mol) W (ml) R (g) C (g) F solution (ml) Initial pH
(g/ml) (mol/mol)
0.10 0.0164 50 0.5 5.00 0.50 0.0096 0.74 7.51
0.10 0.0164 100 0.5 5.00 0.50 0.0048 0.74 7.08
0.10 0.0164 200 0.5 5.00 0.50 0.0024 0.74 6.63
0.10 0.0164 300 0.5 10.00 1.00 0.0032 0.74 6.13
0.10 0.0164 600 0.5 10.00 1.00 0.0016 0.74 6.11
0.20 0.0328 178 0.5 5.00 1.00 0.0054 1.47 –
0.30 0.0492 241 0.5 5.00 1.50 0.0060 2.21 –
0.40 0.0656 292 0.5 5.00 2.00 0.0066 2.95 –
0.50 0.0820 334 0.5 5.00 2.50 0.0072 3.69 –

between 50 and 600 mol/mol (see Table 1). As expected, increasing autocorrelation functions were measured every 3–5 min, only
carbonate concentration (i.e. decreasing R/C ratio) leads to decreas- three representative ones are shown here for clarity.
ing gel time, with an approximately linear dependence of the ob- It can be seen in Fig. 3 that the autocorrelation function
served gel time on the R/C value (i.e. inversely proportional to collected after 10 min of heating has a shape very close to that
the carbonate concentration). We also note that the concentration for an ideal single exponential decay, indicating that the primary
of carbonate (base) causes a systematic variation of solution pH, clusters observed are initially approximately monodisperse and
varying from 7.5 for R/C = 50 mol/mol to 6.1 for R/C = 300 mol/ subject to unhindered Brownian motion. As the time proceeds,
mol (these are initial solution pH values measured at ambient tem- the shape of the autocorrelation function changes in two respects,
perature before solutions were heated), related to neutralisation of as seen in the shape of the autocorrelation function collected after
slightly acidic reactants (pH of the formaldehyde stock solution 75 min of heating: the initial decay rate is lower and there is a sec-
was 4.3 due to small amount of formic acid present, and resor- ondary decay significantly deviating from an ideal monodisperse
cinol–water solution at R/W = 0.1 g/ml had pH of 4.8), see Table 1. exponential decay. This is most likely due to aggregation of pri-
Increasing concentration of resorcinol, corresponding to pro- mary particles forming larger clusters; however, most of them still
portionally increasing concentration of formaldehyde concentra- remain free to move around and appear to be subject to Brownian
tion (as the R/F ratio was kept constant at 0.5 mol/mol), also motion, since the measured autocorrelation function fully decays
caused the gelation time to become shorter, as expected, with an within about 1 ms. Finally, the shape of the autocorrelation func-
approximately linear dependence of the observed gelation time tion measured after 103 min of heating is consistent with one that
on the inverse of resorcinol concentration. In terms of temperature is expected for a gel, consistently with the macroscopic gel time
effect, increasing the temperature by 25 °C (from 55 °C to 80 °C) estimated by tilting a vial containing the gelling solution (see
caused the gelation time to decrease by a factor of about 5, what Fig. 2). At this point, the shape of the autocorrelation function de-
is consistent with expectations and the previous literature [27]. cay shows clear power law features, typical for cases where mobil-
A series of in situ DLS experiments were performed at 55 °C ity of clusters is severely restricted as would be expected for a gel
where the overall gelation process is slow enough to investigate network consisting of a three-dimensional structure of intercon-
formation of primary clusters in more detail. The raw data nected clusters, while unconnected clusters can diffuse within
resulting from these experiments were autocorrelation functions, the pores of the gel structure, as indicated by the initial decay still
which were then fitted to obtain the initial decay rates and from visible.
that the apparent mean hydrodynamic radii of primary clusters A standard analysis of autocorrelation functions can be used to
formed in the solutions. A typical example of how measured auto- estimate the mean hydrodynamic radius of clusters in the solution,
correlation functions change over time is shown in Fig. 3. Although assuming that the clusters are freely diffusing, and they are not yet

Fig. 2. Gelation time as a function of composition for solutions kept at 55 °C (squares) and 80 °C (circles). Lines are guides to eye only. (a) R/W and R/F ratios equal to 0.10 g/
ml and 0.5 mol/mol, respectively, (b) R/C and R/F ratios equal to 100 mol/mol and 0.5 mol/mol, respectively, and sodium carbonate concentration equal to 7.9 mmol/dm3.
K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59 55

Fig. 3. DLS autocorrelation functions (log-linear and log–log plots shown) for solutions with R/C, R/W and R/F ratios equal to 50 mol/mol, 0.10 g/ml and 0.5 mol/mol,
respectively, kept at 55 °C for times as indicated. Dashed lines show theoretical decays calculated assuming a monodisperse cluster population with the hydrodynamic radius
equal to the mean value determined from the initial decay of the autocorrelation function.

interconnected, as evidenced by the autocorrelation function expected to be somewhat larger, so it is clear that primary clusters
following an exponential decay rather than stretched exponential found after the first 10 min are several times larger than single res-
or power law decays characteristic of gel networks. We have there- orcinol–formaldehyde molecules.
fore restricted our attention to the early times where monomodal Further change of the initial decay is very slow, but the second
primary cluster populations are present, so that true mean hydro- decay starts developing at longer times, indicating subsequent
dynamic radii of primary clusters can be determined. The initial aggregation of primary clusters. During this stage, the apparent
decays of measured autocorrelation functions were fitted as de- mean hydrodynamic radius grows only slightly. Eventually, the
scribed in the experimental section, and the resulting initial decay secondary decay increases in relative magnitude and assumes a
rates (C) and the corresponding mean hydrodynamic radii are power law scaling indicative of a gel network (see Figs. 3 and 4).
shown in Fig. 4. Although one could still technically fit the initial decay rate and
Fig. 4b shows clearly that the (true) mean hydrodynamic radius determine the corresponding apparent mean hydrodynamic radius
of primary clusters during the initial stage of their formation (up to during the next stage of the gel formation process, it would not be
about 50 min of heating at 55 °C) is essentially independent of the a physically meaningful quantity describing a real size of growing
R/C ratio: it grows from about 1.5 nm at 10 min to about 3.5 nm at clusters.
50 min, following a power law dependence of size on time. We These observations are consistent with previously published re-
note that the hydrodynamic radius of a hydrated resorcinol mole- sults [15] for experiments performed at 25 °C, although those
cule is about 0.4 nm, which we were able to directly measure in an authors presented the apparent hydrodynamic diameter deter-
aqueous solution of resorcinol using our DLS instrument, while mined by DLS for times extending well into the stage of cluster
variously substituted hydroxymethyl derivatives of resorcinol are aggregation and network formation, where their reported decay

Fig. 4. Time evolution of the initial decay rate (a) and the mean hydrodynamic radius (b) measured at 55 °C. Ratios R/F and R/W were 0.5 mol/mol and 0.10 g/ml, respectively
and R/C as shown in legends.
56 K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59

Fig. 5. DLS autocorrelation functions (log-linear and log–log plots shown) for solutions with R/W and R/F ratios equal to 0.50 g/ml and 0.5 mol/mol, respectively, and sodium
carbonate concentration 7.9 mmol/dm, kept at 55 °C for times as indicated.

time spectrum became bimodal and very long decay times were times where decays were still following a single exponential pat-
observed. However, in the early stage where monomodal decay tern, and so, there was no hindrance on free diffusion of primary
spectra were observed (see their Fig. 1), initial sizes of primary clusters (in the same manner as it was done for samples with var-
clusters appear to be quite similar for various values of the R/C ra- ied R/C ratios). The results of this analysis are shown in Fig. 6. One
tio, although there are only few data points reported for hydrody- can clearly see that the mean hydrodynamic radius is again very
namic diameters below 5 nm (see their Fig. 2). Similarly, the similar for all values of R/W ratios investigated here, indicating
growth of primary clusters was investigated at 25 °C by SAXS that the overall concentration of reactants does not control the
[17], and from the data reported in that work, it can be seen that sizes of primary clusters. This is also consistent with previous
the radius of gyration of primary clusters appears to be around observations of Yamamoto et al. [15] at 25 °C, where reported
2 nm (see their Fig. 4) for all conditions investigated at the same hydrodynamic diameters of primary clusters were similar for a
time of 500 min, which is well before any increase in viscosity indi- range of R/W ratios at early times (see their Fig. 2), while they be-
cating approaching gelation. The same authors also showed scaling came very different at later times (where these can only be seen as
exponents for their SAXS intensity data increasing over time from apparent hydrodynamic diameters which do not reflect real cluster
their initial values around 2 to plateau at values around 4 at longer sizes anymore, as discussed above).
times. The power law scaling exponent of 4 is indicative of smooth From the data reported in Figs. 4 and 6 above, we can see that
surfaces (Porod regime), while value below 3 is considered to be the size of primary clusters does not appear to vary significantly
indicative of mass fractal structures. However, in the reported data with the concentration of carbonate or reactants. This might seem
(see their Fig. 3), there is an extremely narrow range of q values surprising, because carbonate is a catalyst for the substitution of
over which a power law scaling is fitted, and it can be argued that resorcinol with formaldehyde, and therefore, it is expected that
such data cannot be used reliably to extract structural information there is more resorcinol–formaldehyde present and hence faster
from scattered intensities. In fact, it has been shown previously polymerisation at higher carbonate concentration. However, there
that for small fractal-like aggregates, the apparent power law scal- is a clear effect of the carbonate concentration as well as the reac-
ing determined from scattering structure factors can be signifi- tants concentration on the measured values of scattered intensities
cantly lower than the actual mass fractal scaling exponent and as shown in Fig. 7.
only approached the actual value when aggregates are sufficiently In Fig. 7a, we can see that scattered intensities (expressed here
large [28]. in terms of mean count rates measured at the scattering angle of
A series of experiments with varied R/W ratios were also per- 90°) stay approximately constant during the first 5–10 min and
formed at 55 °C in order to determine the influence of overall reac- then start increasing steadily over time. However, unlike mean
tants concentration on the evolution of autocorrelation function hydrodynamic radii, there is a clear dependence of scattered inten-
and the apparent hydrodynamic radius. Ratio of resorcinol to form- sities on the R/C ratio, with higher intensities for smaller R/C ratios
aldehyde was kept at 0.5 mol/mol, and instead of keeping R/C ratio (i.e. higher carbonate concentrations). Since hydrodynamic radii
constant, the concentration of sodium carbonate was kept constant were very similar at all three R/C ratios (see Fig. 4), while the
at and equal to 7.9 mmol/dm, while the ratios of resorcinol to scattered intensity is higher for higher carbonate concentration
water were changing from 0.10 g/ml to 0.50 g/ml. Fig. 5 shows typ- at a given time, it follows that the scatterers (i.e. primary clusters)
ical autocorrelation functions for a sample with R/W 0.50 g/ml. are either present at higher concentrations or they possess higher
Interestingly, the measured autocorrelation functions are very sim- optical contrast (difference in refractive index respective to sur-
ilar to those shown in Fig. 3, and the decay times are comparable, rounding solvent) at higher carbonate concentrations. Since the
despite the fivefold increase in the resorcinol concentration (cf. surrounding solvent is very similar in all cases (same reactant con-
curves for 10 min in Figs. 3 and 5). centrations in water, with carbonate being present at comparably
Mean hydrodynamic radii were calculated from the autocorre- low concentrations), it is most likely that in the case of various
lation functions for samples with varied R/W ratios for a range of R/C ratios, the differences in scattered intensity are due to more
K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59 57

Fig. 6. Time evolution of the initial decay rate (a) and the mean hydrodynamic radius (b) at 55 °C. Ratio R/F was 0.5 mol/mol, sodium carbonate concentration was 7.9 mmol/
dm3, and R/W was as shown in the legend.

primary clusters formed at higher carbonate concentrations. This is estimated to vary between 1.35 for R/W = 0.1 g/ml and 1.40 for
reasonable, since higher carbonate concentration should imply fas- R/W = 0.5 g/ml [29], while the typical literature values for dense li-
ter substitution reaction and so faster production of substituted quid phase systems chemically similar to the clusters of reaction
resorcinol and presumably also faster subsequent condensation intermediates have been reported between 1.44 (saturated aque-
reaction. ous solution of phenol; [30]) and 1.48 (phenol–formaldehyde resin
In Fig. 7b, we can see that scattered intensities are lower for at 50% w/w solids; [31]). Therefore, it can be expected that primary
higher concentrations of resorcinol (and formaldehyde, since the clusters composed of resorcinol–formaldehyde reaction intermedi-
R/F ratio was kept constant equal to 0.5). By the same argument ates have higher optical density than surrounding bulk solution,
as above, the primary clusters are either present at lower concen- and thus indeed, their optical contrast will be lower in solutions
trations or they have lower optical contrast. Since it would be ex- with higher reactant concentrations, i.e. at higher R/W ratios,
pected that the primary clusters would become more (not less) which have higher refractive index.
numerous at higher concentration of reactants (while the carbon- A series of DLS experiments were also performed on solutions
ate concentration was held constant at all R/W ratios), it follows reacting at 80 °C, while quenching them at specified times to
that they ought to have decreasing optical contrast with increasing 20 °C where reactions are expected to be much slower and DLS mea-
reactant concentration. This is sensible, since composition of surements could be performed. An example of results from these
solvent surrounding primary clusters becomes more similar to experiments is shown in Fig. 8. Comparison of Figs. 3 and 8 shows
clusters as concentration of reactants increases and water concen- that the evolution of the autocorrelation function indicates the
tration decreases in the reaction mixture with increasing R/W ra- same overall pattern at both temperatures – from an initially mono-
tios. Refractive indices of reaction mixtures used here were disperse population of primary clusters, through a polydisperse

Fig. 7. Time evolution of the scattered intensity measured at 55 °C. (a) Ratios R/F and R/W were 0.5 mol/mol and 0.10 g/ml, respectively, and R/C was as shown in the legend.
(b) Ratio R/F was 0.5 mol/mol, sodium carbonate concentration was 7.9 mmol/dm3, and R/W was as shown in the legend.
58 K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59

Fig. 8. DLS autocorrelation functions (log-linear and log–log plots shown) for solutions with R/C, R/W and R/F ratios equal to 200 mol/mol, 0.10 g/ml and 0.5 mol/mol,
respectively, kept at 80 °C for times as indicated and then rapidly quenched and measured at 20 °C.

collection of aggregated clusters to a gel network of interconnected resorcinol or its oligomers) reach a limit of their solubility/misci-
clusters. bility in the solution mixture at a given temperature, and the
However, there is a significant difference in time when the solution phase undergoes a process similar to spontaneous nano-
autocorrelation function falls into one of these three stages. For emulsification or micelle formation. This leads to primary clusters
example, at the composition corresponding to results shown in observed in this work as well as by others using in situ DLS or SAXS
Fig. 3, the aggregation stage was observed only after more than measurements (see above). Such a process is controlled by solubil-
50 min at 55 °C, while it appeared in just 10 min at 80 °C. Similarly, ity/miscibility equilibria with demixing resulting in formation of
the solution gelled about five times later at 55 °C than at 80 °C (see discrete non-interconnected primary clusters freely diffusing in
Fig. 3). These results indicate that the temperature influences the surrounding solution matrix, in contrast to two interpenetrating
overall rate of the gel formation, but it does not appear to influence phases resulting from a classical polymerisation driven microphase
an overall mechanism of its formation, at least within the range of separation.
temperatures investigated. As indicated in the overall reaction mechanism shown in Fig. 1,
In order to assess the effect of temperature on the mean hydro- there are a series of reactions starting with the first resorcinol sub-
dynamic radius of primary clusters, in Fig. 9, we show a compari- stitution with formaldehyde, followed by the second and possibly
son of mean hydrodynamic radii of primary clusters obtained by further ones. Then, the substituted resorcinol species undergo sub-
heating at 55 °C and 80 °C. It can be that the primary clusters ob- sequent condensation reactions to form dimers and further oligo-
tained at 80 °C are larger than those at 55 °C, while the time evolu- mers. It is likely that a certain substituted and/or condensed
tion is following the same overall trend irrespective of the R/C species has limited solubility/miscibility in the background reac-
ratio. tion mixture, and as it accumulates over time, it reaches a critical
We have seen that under the range of conditions investigated concentration at which primary clusters are formed. As can be seen
here, the size of primary clusters varies only little when changing in Fig. 7a, there is an earlier onset of an increase in scattered inten-
concentrations of reactants or carbonate at a given temperature. sity at higher carbonate concentrations indicating that such a crit-
Furthermore, primary clusters are more numerous, and overall ical concentration of reaction intermediates is reached earlier, as
gelation process is faster at higher carbonate concentrations. If pri-
mary clusters are more numerous, then it is sensible that their
aggregation is faster, and space-filling by growing aggregates leads
to gel formation at shorter times.
It may seem intriguing that the primary cluster size is only little
sensitive to the reaction mixture composition. This appears to be
inconsistent with the scenario where growth of primary clusters
is driven by homogeneous polymerisation of substituted resor-
cinol, as proposed previously [17], since faster reactions should
lead to faster growth of cluster size, but this is not the case at early
stage of the process before the onset of primary cluster aggrega-
tion. However, our results are consistent with a process of nano-
scale demixing, controlled by solubility/miscibility equilibria,
similar to that involved in micellisation or spontaneous emulsifica-
tion, such as in Ouzo effect. Unlike the classical Ouzo effect in
non-reacting mixtures, though, resorcinol–formaldehyde species
in primary clusters undergo further reactions, and hence, chemical
and physical nature of clusters evolves over time. Fig. 9. Early evolution of mean hydrodynamic radius for solutions kept at 55 °C
We propose that certain intermediates produced in resorcinol– (squares) and 80 °C (circles). (R/C ratios 50, 100 and 200 mol/mol 1; R/F is 0.5 mol/
formaldehyde polymerisation (either multiply substituted mol 1; R/W is 0.10 g/ml 1).
K.Z. Gaca, J. Sefcik / Journal of Colloid and Interface Science 406 (2013) 51–59 59

expected for a series of reactions catalysed by bases, such as car- intermediates. This leads to a competition between polymerisa-
bonates. At higher carbonate or reactant concentrations, reaction tion, phase separation, coalescence and aggregation resulting in
rates are higher, more intermediates are produced, and therefore, nanoscale morphologies of wet gels. Analogous formation mecha-
more primary clusters (of approximately the same size) are formed nisms have been previously proposed for several polymerisation
within the same solution matrix, eventually resulting in faster and sol–gel systems, including monodisperse silica, organosilicates
aggregation and gelation. Since intermediate species in primary and zeolites.
clusters undergo further reactions, chemical and physical nature There are likely to be multiple roles played by carbonates in the
of clusters continually evolves, which may help to explain why process of resorcinol–formaldehyde gelation. In addition to car-
their size is changing over time even though their formation is bonates being catalysts for reactions of resorcinol and formalde-
thermodynamically controlled. Primary clusters may be initially li- hyde, resulting in faster reactions and higher numbers of primary
quid-like and subject to subsequent coalescence as well as aggre- clusters formed, they may also influence solubility/miscibility
gation. Further polymerisation may proceed faster within the equilibria in complex multicomponent reaction mixtures responsi-
clusters than in the bulk solution due to locally higher concentra- ble for formation of primary clusters. Furthermore, carbonates may
tions of intermediates, eventually leading to solidification and also influence colloidal stability of primary clusters resulting in
colloidal destabilisation or polymerisation driven microphase faster aggregation and subsequent gelation at higher carbonate
separation. concentrations.
The final size of particles observed to form the gel network then
results from a delicate interplay of polymerisation, phase equilib- References
ria, coalescence and aggregation phenomena. Furthermore, car-
bonates might play role in colloidal stability of primary clusters [1] R.W. Pekala, J. Mater. Sci. 24 (1989) 3221.
[2] A.C. Pierre, G.M. Pajonk, Chem. Rev. 102 (2002) 4243.
or their aggregates, resulting in faster aggregation and subsequent [3] S.A. Al-Muhtaseb, J.A. Ritter, Adv. Mater. 15 (2003) 101.
gelation at higher carbonate concentrations, which could also con- [4] M. Mirzaeian, P.J. Hall, Electrochim. Acta 54 (2009) 7444.
tribute to the observed effect of carbonates on gel times. [5] M. Mirzaeian, P.J. Hall, J. Mater. Sci. 44 (2009) 2705.
[6] H. Kabbour, T.F. Baumann, J.H. Satcher, A. Saulnier, C.C. Ahn, Chem. Mater. 18
(2006) 6085.
4. Conclusions [7] N. Tonanon, Y. Wareenin, A. Siyasukh, W. Tanthapanichakoon, H. Nishihara,
S.R. Mukai, H. Tamon, J. Non-Cryst. Solids 352 (2006) 5683.
[8] A.C. Pierre, in: Kirk-Othmer Encyclopedia of Chemical Technology, John Wiley
Formation of primary clusters and their subsequent aggregation & Sons Inc., New York, 2006.
and gelation during resorcinol formaldehyde polymerisation was [9] R.B. Durairaj, Resorcinol: Chemistry, Technology and Applications, Springer,
monitored by DLS at two different temperatures, 55 °C and 80 °C. Berlin, 2005.
[10] C. Lin, J.A. Ritter, Carbon 35 (1997) 1271.
We found that the kinetics of growth of primary clusters, before [11] N. Job, C.J. Gommes, R. Pirard, J.-P. Pirard, J. Non-Cryst. Solids 354 (2008) 4698.
they were subject to further aggregation, changed only very little [12] H.R. Gerberich, G.C. Seaman, Formaldehyde, in: Kirk-Othmer Encyclopedia of
when the carbonate concentration was varied. Although more pri- Chemical Technology, John Wiley & Sons, Inc., 2008.
[13] R.W. Pekala, D.W. Schaefer, Macromolecules 26 (1993) 5487.
mary clusters were formed at higher carbonate concentrations, as [14] D.W. Schaefer, G. Beaucage, R.W. Pekala, J. Non-Cryst. Solids 186 (1995) 159.
deduced from scattered light intensities, and corresponding gela- [15] T. Yamamoto, T. Yoshida, T. Suzuki, S.R. Mukai, H. Tamon, J. Colloid Interface
tion times were shorter, the size of primary clusters was very sim- Sci. 245 (2002) 391.
[16] C.J. Gommes, A.P. Roberts, Phys. Rev. E 77 (2008).
ilar at all carbonate concentrations. We also found that the growth [17] H. Tamon, H. Ishizaka, J. Colloid Interface Sci. 206 (1998) 577.
kinetics of primary clusters was only little dependent as the con- [18] O. Czakkel, A. Madsen, Europhys. Lett. 95 (2011).
centration of reactants (resorcinol and formaldehyde) was varied. [19] J. Sefcik, A.V. McCormick, Catal. Today 35 (1997) 205.
[20] D.L. Green, S. Jayasundara, Y.-F. Lam, M.T. Harris, J. Non-Cryst. Solids 315
Resulting gelation times were again shorter at higher reactant con- (2003) 166.
centrations, indicating that more primary clusters were produced, [21] D.D. Kragten, J.M. Fedeyko, K.R. Sawant, J.D. Rimer, D.G. Vlachos, R.F. Lobo, M.
although this could not be unequivocally established from scatter- Tsapatsis, J. Phys. Chem. B 107 (2003) 10006.
[22] S.E. Rankin, J. Sefcik, A.V. McCormick, J. Phys. Chem. A 103 (1999) 4233.
ing intensities.
[23] J. Ambati, S.E. Rankin, J. Colloid Interface Sci. 362 (2011) 345.
Our results indicate that size of primary clusters appears to be [24] C.J. Gommes, N. Job, J.-P. Pirard, S. Blacher, B. Goderis, J. Appl. Crystallogr. 41
thermodynamically controlled, where a miscibility limit is reached (2008) 663.
due to formation of certain reaction intermediates, most likely [25] K.Z. Gaca, in: Kinetics and Mechanisms of Early Stages of Resorcinol–
Formaldehyde Polymerization; PhD thesis, University of Strathclyde,
with relatively small molecular weight, resulting in nanoscale Glasgow, United Kingdom, 2012.
molecular demixing leading to formation of approximately mono- [26] R. Finsy, Adv. Colloid Interface Sci. 52 (1994) 79.
disperse primary clusters, similar to formation of micelles or spon- [27] N. Job, F. Panariello, M. Crine, J.-P. Pirard, A. Léonard, Colloids Surf. A 293
(2007) 224.
taneous nanoemulsions. At higher carbonate concentration or [28] P. Sandkuhler, J. Sefcik, M. Morbidelli, Langmuir 21 (2005) 2062.
reactant concentrations, more primary clusters were produced [29] A.K. Jain, Eur. J. Pharm. Biopharm. 68 (2008) 701.
due to faster reactions. Primary clusters may be initially liquid-like, [30] H. Ghanadzadeh Gilani, A. Ghanadzadeh Gilani, M. Sangashekana, J. Chem.
Thermodyn. 58 (2013) 142.
and further polymerisation may proceed faster within them than [31] J. Monni, L. Alvila, T.T. Pakkanen, Ind. Eng. Chem. Res. 46 (2007) 6916.
in the bulk solution due to locally higher concentrations of

You might also like