You are on page 1of 13

European Polymer Journal 132 (2020) 109719

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

Temperature effect on aqueous NH4CN polymerization: Relationship T


between kinetic behaviour and structural properties

Itziar Masa, José L. de la Fuenteb, Marta Ruiz-Bermejoa,
a
Centro de Astrobiología (INTA-CSIC), Dpto. Evolución Molecular, Ctra. Torrejón-Ajalvir, km 4, Torrejón de Ardoz, 28850 Madrid, Spain
b
Instituto Nacional de Técnica Aeroespacial “Esteban Terradas” (INTA), Ctra. Torrejón-Ajalvir, km 4, Torrejón de Ardoz, 28850 Madrid, Spain

A R T I C LE I N FO A B S T R A C T

Keywords: Herein, a kinetic analysis for aqueous NH4CN polymerizations is presented, which is consistent with an auto-
HCN polymers catalytic model when polymerizations are performed at relatively high temperatures, 80–90 °C. Further ex-
Precipitation polymerization periments at lower polymerization temperatures, approximately 50 °C, have demonstrated that this relevant
Kamal model prebiotic reaction follows nth-order kinetics rather than an autocatalytic mechanism. In addition, the sol frac-
Autocatalysis
tions of these precipitation polymerizations have been evaluated by UV–Vis measurements, which also show a
Kinetic
Spectroscopy
mechanistic shift with the reaction temperature. This change in the kinetic behaviour led to the proposal of a
simple empirical methodology to describe both chemical- and diffusion-controlled regions. Despite the simplicity
of the approach based on the Hill equation, fundamental kinetic parameters, such as the activation energy, can
be determined in the diffusion-free zone. These results motivated a systematic structural characterization study
of the respective insoluble polymers by means of elemental analysis, FT-IR and NMR spectroscopies and XRD. All
these kinetic and structural analyses confirmed that the temperature has a significant effect on the poly-
merization kinetic of the system, on the macrostructural features and properties of the HCN-based polymers, and
presumably also on the polymerization pathways. These data increase our knowledge about the chemistry of this
particular family of HCN polymers, which is currently of interest both in the field of materials science and in
prebiotic chemistry and astrobiology.

1. Introduction initial pH, the salinity of the aqueous environment, temperature, pre-
sence of oxygen, or the reaction time [see e.g. 2, 9–12]. Therefore, a
Currently, the development of new and advanced functional mate- comprehensive knowledge of the kinetic history of HCN polymer for-
rials must be considered from an environmental perspective. This mation and their respective structural features are prerequisites to
starting point implies the challenge of finding sustainable and green properly tune the properties of the desired reaction products. Thus, in
synthetic methods. The use of solvents such as water and low reaction the present case, the effect of the temperature is systematically studied
temperatures are fundamental conditions to obtain efficient, friendly for aqueous NH4CN polymerization considering several approaches and
and low-cost industrial processes. In this sense, the production of HCN- models to explain the kinetics of this precipitating reaction as well as its
derived polymers might be a good example, because this heterogeneous influence on the structural nature of the insoluble polymers synthe-
family of substances can be obtained under a wide range of experi- sized.
mental conditions, e.g., using water as a solvent in spontaneous poly- In a previous study, a Kamal autocatalysis kinetic model was pro-
merization reactions, or by bulk polymerizations at not relatively high posed to describe the formation of insoluble NH4CN polymers and gel
temperatures [1]. These complex macromolecular systems show inter- fractions during the polymerization of NH4CN in aqueous medium at
esting properties [2], leading to their consideration as promising mul- relatively high temperatures, 75–90 °C [12]. Under those conditions,
tifunctional materials with potential applications as coating materials, the maximum reaction rate occurred at low conversions, and an auto-
semiconductors, photocatalysts or capacitors (see, e.g., [2–8]). How- catalytic kinetic model for these precipitation polymerizations must be
ever, their structural characteristics and, by extension, their final considered. According to the behaviour of an autocatalytic poly-
properties and hence their potential applications depend directly on the merization, the equation dα/dt = k αm (1-α)n, where m and n are re-
experimental synthetic conditions used for their production, such as action orders and k is the reaction rate constant, could be used.


Corresponding author.
E-mail address: ruizbm@cab.inta-csic.es (M. Ruiz-Bermejo).

https://doi.org/10.1016/j.eurpolymj.2020.109719
Received 27 February 2020; Received in revised form 20 April 2020; Accepted 21 April 2020
Available online 01 May 2020
0014-3057/ © 2020 Elsevier Ltd. All rights reserved.
I. Mas, et al. European Polymer Journal 132 (2020) 109719

However, due to the very complex pathway proposed for the cyanide considered primary substances in prebiotic chemistry and astrobiology.
polymerization [2], where several events occur simultaneously, kinetic
models with multiple rates can provide more accurate results. There- 2. Experimental section
fore, the well-known Kamal-Sourour kinetic model was chosen. This
model involves two rate constants, k1 and k2 (with two different acti- 2.1. Aqueous NH4CN polymerizations
vation energies and pre-exponential factors), and has the following
form, dα/dt = (k1 + k2 αm)·(1-α)n, where n and m are the reaction NH4CN polymers were synthesized from equimolar water solutions
orders. In this case, the influence of the reaction products on the re- of NaCN and NH4Cl (1 M) under atmospheric pressure. The poly-
action rate is given by the term k2αm. The application of this phenom- merization reactions were performed in 20 mL glass vials fitted with
enological kinetic model instead of mechanistic models was based on Teflon/silicone septa and placed into heated metal blocks with tem-
two relevant aspects: on the one hand, the simplicity of the phenom- peratures of 50, 60 and 70 °C as well as 80 and 90 °C, as comparative
enological models, and on the other hand, as was indicated above, the experiments at higher temperatures, finding a good correlation between
extraordinary complexity of the aqueous heterogeneous NH4CN poly- the previous data [12] and those reported herein (Fig. S1). When the
merization, which is convoluted with hydrolysis, oxidation and other desired time was achieved, the reaction mixture was immediately
thermally initiated chemical reactions [2]. However, certain limitations cooled, and the final suspensions were filtered. The reaction products
were found when this phenomenological approach was employed. Low consisted of both soluble and insoluble materials. Insoluble black solids
correlation coefficients were obtained, especially at lower polymeriza- and brown solutions (gel and sol fractions, respectively) were collected,
tion temperatures, and unrealistic reaction orders were determined. In and both were freeze-dried until a constant weight was reached, re-
addition, an analysis of the activation energy (Ea) was only performed sulting in black powdered solids (insoluble polymers) and brown solids
from three temperatures in a narrow range, and then, a reliable de- (soluble oligomers/polymers). Polymer insoluble conversions, α, were
termination of the Arrhenius parameters was not achieved. These re- determined by a gravimetric method and were calculated as the initial
sults indicated a change in the model with the polymerization tem- weight of the cyanide added to the system [α = (final weight of in-
perature and/or with the conversion range under study. Thus, herein, soluble NH4CN polymer/initial weight of CN−) × 100].
new experimental data sets have been presented for medium tem- The insoluble black polymers were characterized by elemental
peratures, 50–70 °C, and replicated for 80 and 90° C in order to solve analysis, FT-IR spectroscopy, 13C NMR spectroscopy, and X-ray dif-
some of these considerations. fraction. Additionally, the acid-hydrolysed products of these gel frac-
Moreover, aqueous NH4CN polymerizations at lower temperatures, tions were characterized by elemental analysis, FT-IR spectroscopy and
below 38 °C, were previously monitored by UV–Vis spectroscopy in an UV–Vis spectroscopy.
attempt to model the kinetics of NH4CN polymer formation through the On the other hand, the lyophilized sol fractions were analysed in a
spectroscopic characteristics of the sol fractions [10]. In that case, the similar way to previously described for the crude reactions of the
Hill equation, y = Vmax [xn/(kn + xn)], where n is the Hill coefficient NH4CN polymerizations below 38 °C by UV–Vis spectroscopy [10]. In
(unitless) that reflects the steepness (sigmoidicity) of the curve, was the present case, 4 mg of the dried sol samples were dissolved in 1 mL of
used taking in consideration the evolution of the absorbance relation- water. The suspensions were centrifuged at 13000 rpm for 10 min, and
ships at A230/A300 during the polymerization process for the sol frac- the UV–Vis spectra of the supernatants were recorded.
tions. The band at ~300 nm decreases in a time-dependent manner up
its disappearance, which indicates the transformation of cyanide and/or 2.2. Elemental analysis
the tetramer diaminomaleonitrile (DAMN, likely the main intermediate
in the HCN or cyanide polymerization, as will be explained later) in The NH4CN polymers were examined for the determination of the
other products. In parallel, the narrow peak at ~230 nm increases with mass fractions of carbon, hydrogen and nitrogen using a Perkin Elmer
time, thus indicating the reaction polymerization progress. The Hill elemental analyser, model CNHS-2400. The percentages of oxygen in
equation is widely used for the description of biochemical, geochemical the samples were obtained by calculating the differences.
and chemical processes (for more detail about the Hill function, see the
references cited in [10]). This simple model was very useful to provide 2.3. FT-IR spectroscopy
a general view of the influence of the air, ammonium cation and salty
environment in the kinetic process of NH4CN polymerization at 20 and Diffuse reflectance spectra were acquired in the 4000–400 cm−1
38 °C. Here, this experimental modelling for the sol fractions is ex- spectral region using an FT-IR spectrometer (Nicolet, model NEXUS
tended to a range of temperatures from 50 to 90 °C for NH4CN polymers 670) configured with a drift reflectance accessory (Harrick, model
as a complementary study of the gravimetric measurements and kinetic Praying Mantis DRP) mounted inside the instrument compartment. The
analysis from the gel fractions indicated above. spectra were obtained in CsI pellets, and the spectral resolution was
On the other hand, the black insoluble solids obtained in the water 2 cm−1.
polymerization of NH4CN from 50 to 90 °C were characterized using
13
several techniques, such as elemental analysis, FT-IR spectroscopy, 2.4. C CP MAS-NMR
solid-state 13C NMR spectroscopy, and XRD. In addition, the stability of
13
these solids against acid hydrolysis was studied. All this structural in- C CP MAS-NMR spectra were obtained using a Bruker Advance
formation may contribute to explaining the complex pathways of pro- 400 spectrometer and a standard cross-polarization pulse sequence.
duction of HCN polymers and to relating the changes in the kinetic and Samples were spun at 10 kHz, and the spectrometer frequency was set
mechanistic approaches described herein with the main plausible in- to 100.62 MHz. A contact time of 1 ms and a period between successive
termediate molecules implicated in the formation of this relevant accumulations of 5 s were used. The number of scans was 5000, and the
polymeric system. The production of HCN polymers in water environ- chemical shift values were referenced to tetramethylsilane (TMS).
ments is known to be a robust process. The results presented here and
those recently reported in the literature corroborate this fact as well as 2.5. Powder X-ray diffraction
the direct dependence of the properties of the final polymerization
products on the experimental conditions of the reaction. These con- Powder X-ray diffraction was performed using a Bruker D8 Eco
clusions deepen and extend our vision about the nature and chemistry Advance with Cu Kα radiation (λ = 1.542 Å) and a Lynxeye XE-T linear
of these fascinating polymeric systems, known as HCN polymers, which detector. The X-ray generator was set to an acceleration voltage of
have recently been of high interest in materials science but are also 40 kV and a filament emission of 25 mA. Samples were scanned

2
I. Mas, et al. European Polymer Journal 132 (2020) 109719

between 5° (2θ) and 60° (2θ) using a step size of 0.05° and a count time
of 1 s, using the Bragg–Brentano geometry.

2.6. UV–Vis spectroscopy

The UV–Vis spectra were recorded in a Nanodrop 1000 spectro-


photometer (Thermo Scientific, Madison, Wisconsin, EE. UU) in the
range of 220 to 730 nm at room temperature. To obtain reliable spec-
troscopic data, it was necessary to pre-process the raw spectra before
the kinetic analysis. Thus, the spectra were normalized and then
smoothed by applying the Savitzky-Golay algorithm.

2.7. Hydrolysis conditions

Twenty milligrams of selected black insoluble polymers at limit


conversions were hydrolysed in flame-sealed glass vials using 0.5 mL of
6 M HCl at 110 °C for 24 h. The crude reactions were centrifuged and
washed with water (13000 rpm/10 min and 0.5 mL of water × 3). The
pellets and the supernatants were collected and freeze-dried until a
constant weight was achieved.

3. Results and discussion

3.1. Polymerization kinetics

3.1.1. Analysis of the production of insoluble NH4CN polymers by


gravimetry
The isothermal polymerization kinetics of NH4CN were system-
atically investigated at five different temperatures between 50 and
90 °C at a 10 °C interval. Fig. 1a) shows the plot of the experimentally
determined conversion (α) versus polymerization time (t) for all tem-
peratures. As can be seen, the percentage conversion increased rapidly
with the reaction time at the initial stage, followed by the next stage,
where the rate of increase decreased asymptotically thereafter until the
end of the polymerization process. This general characteristic beha-
viour of NH4CN polymerization in water can be described by the Hill
equation [12], as it has been determined (lines curves). This proposal is
based on a previous work [10], where the sol fractions during this Fig. 1. a) Degree of conversion α and b) the conversion rate dα/dt vs reaction
polymerization process were analysed by means of spectroscopic mea- time of aqueous NH4CN polymerization at different isothermal temperatures.
surements as indicated in the Introduction. It is important to note that Note: The insets have the same axis legends as that of the full figure.
excellent correlation coefficient values (R2), higher than 0.98, were
found, which are collected in Table S1, together with the parameter of where the reaction changes from a chemically controlled to diffusion-
the Hill equation. Although, in principle, some of these parameters have controlled process. In Fig. 1a), a solid black line showing the division of
no physical meaning, it is significant that the n exponent decreases as chemically controlled and diffusion-controlled regions is drawn. The
the polymerization temperature does, except for the temperature of αonset values for all the polymerizations under study are also presented
50 °C. in Table S1, taking a constant value of approximately 25% for the
The Hill function gives a correct fit to all experimental set data, and higher temperatures and after increasing as the polymerization tem-
the profile of these curves indicates that diffusion effects occur when perature decreases.
the reaction progresses, as can be observed in Fig. 1a). From these An advantage of this approach is its simplicity in presentation,
curves, it is possible to build an empirical model [13], which requires though there may be some inconsistency in the transition region be-
much less overall investment in modelling when little is known about tween both stages. In contrast, the current empirical model can be very
the physical and chemical phenomena underlying the process, as occurs useful in the early stage of polymerization, as observed in the plot inset.
in this case. Therefore, at the initial regime or stage, where the reaction Thus, the fitting curves of α-t conform to an apparent “self-accelera-
kinetics are likely chemically controlled, the instantaneous poly- tion”, S shape, for the lower temperatures, especially at 50 °C. However,
merization rate (Rp) can be determined through the slope of the straight this profile for the autocatalysis process is not as evident for 80 and
lines drawn (dashed lines). The obtained values were used later for the 90 °C. This preliminary idea must be confirmed through phenomen-
activation energy (Ea) analysis. The other distinctive parameter of the ological kinetic models; consequently, new attempts to fit the data were
second kinetic stage is the limit conversion (α∞), which was determined made in order to clarify the uncertainties mentioned regarding the
from the last region, where the curves are asymptotic to the time-axis applicability of the Kamal model.
(dashed lines), and these values are compiled in Table S1. Here, a de- Before this analysis, it is important to note that in general, the ki-
crease in the reaction species’ mobility occurs, and the polymerization netic models of thermosetting polymers fall into two main types: (1)
reaction becomes diffusion-controlled because of an increase in the nth-order models and (2) autocatalytic models. This classification is
molecular weight and/or the crosslinking of the polymer chains. The based on the fact that there are two categories of crosslinking reactions
α∞ values vary between 35 and 44% conversion, reaching the highest differentiated by whether the maximum rate of reactions takes place at
plateau conversion at lower reaction temperatures. The crossing point zero conversion or at a finite conversion value. The nth-order model is
of both lines can give an estimation of the time of conversion (αonset)

3
I. Mas, et al. European Polymer Journal 132 (2020) 109719

often simply expressed by the reaction rate equation, dα/dt = k(1-α)n, used. In this case, k1 ≤ 0.001 and k2 and m are very low; therefore, the
where n is the reaction order, dα/dt is the reaction rate, and k is the autocatalytic effect is very weak and can be considered negligible. The
apparent rate constant. This equation assumes that n is a constant in- polymerization reaction resembles an nth-order process, but the appli-
dependent of temperature and conversion a predicts a linear relation- cation of this kinetic model does not improve the correlation coefficient
ship between ln(dα/dt) and ln(1-α). In most works so far, these as- when a nonlinear multiple regression method is used. Nevertheless, a
sumptions were accepted, though there are often large discrepancies significantly better regression coefficient is obtained when a lineariza-
from reality. However, the nth-order model cannot describe the pro- tion of the nth-order equation is applied. This fact can be observed
gress of those systems in which autocatalysis is present, for those cases through the plot inset in Fig. 2, where ln(dα/dt) versus ln(1-α) is pre-
the well-known Kamal model can be used. sented for these temperatures. The polymerization kinetics at 50 °C
From the experimental data of conversion as a function of poly- could not be fitted properly to the Kamal model, but the nth-order
merization time, it is possible to obtain the reaction rate, dα/dt, by equation can correctly predict the experimental data, as illustrated in
differentiating the data sets for the α-t curves, as shown in Fig. 1b). In this figure (dashed lines). However, it is observed that the opposite
the inset plot of this figure, it is possible to see in detail the profiles of behaviour occurs at this lowest temperature. In this case, the relation-
the new curves for the first stage of the polymerizations. An unclear ship between ln(dα/dt) and ln(1-α) presents a low linearity, though a
tendency is found for the derivatized experimental data at poly- better correlation coefficient, R2 > 0.96, is found when the other
merization temperatures of 90 and 80 °C; however, the conversion rate method of fitting is used.
data at lower temperatures present less dispersion, as can be observed All these results for both kinetic models are summarized in Table S2
in the new plot inset. In addition, these temperatures of 70, 60 and for the three temperatures under study. In any case, the results indicate
50 °C have not been previously studied, and therefore, an exhaustive that the reaction order “n” may not be constant throughout the range of
analysis must be completed. Regardless of the difficulties in following temperatures. In addition, these results reflect a possible change in the
the polymerization at the fast-early stage using gravimetric measure- mechanism. At higher temperatures, the polymerization can follow the
ments, the results for 70 °C indicate that the conversion rate increased autocatalytic kinetic model rather than an nth-order kinetic model.
rapidly at the start of the reaction to a maximum and then decreased
exponentially to a minimum value. This behaviour, typical of an au-
tocatalytic process, is not as obvious for the data set at 60 °C, and even 3.1.2. Activation energy analysis
for 50 °C, where the maximum conversion rate appears to occur at t = 0 Consistent estimation of the Arrhenius parameters requires one to
or α = 0, in which case the nth-order model could be considered. conduct experiments at three or more temperatures. In addition, typi-
Attempts were made to curve fit the dα/dt data – α plots, as shown cally, the temperature range should not be less than 20–40 °C. However,
in Fig. 2, using the Kamal expression for the obtained values from the certain limitations found in a preliminary work led us to calculate this
polymerization temperatures of 70 and 60 °C. The kinetics parameters relevant kinetic parameter with only three temperatures and in a
k1, k2, m and n from this curve fitting are given in Table S2. These narrow temperature range from 75 to 90 °C [12]. Although the new
kinetic parameters were calculated by using multi-variable nonlinear experiments presented in this work covered a rather wide temperature
regression, and the square of the coefficient of correlation was also range, the polymerization reaction under study seems to be dominated
collected in this table. It can be seen from this figure (solid lines) and by a dissimilar mechanism at different stages of the process and, at the
Table S2 that a good correlation was only found for 70 °C. Although the same time, involves an autocatalytic mechanism operating at higher
error of fitting was acceptable considering experimental uncertainties temperatures that shifts to an nth-order reaction as the polymerization
and the fact that a reduced number of kinetic data sets are fitted to the temperature is reduced.
phenomenological kinetic model, these kinetic parameters were in This fact makes it complicated to address the Ea determination, and
concordance with those reported for our group at higher temperatures, therefore, this determination has been done from the values of reaction
90 and 80 °C (see Table 1 in reference [12]). Thus, the k2 values are rate obtained from the first polymerization stage, which is chemically
much higher than k1, and at the same time, m ≈ 1.5 and n take a high, controlled. The usual Arrhenius dependence constructed from the Rp
unrealistic value. In contrast, the statistical parameter R2 was not good values in Table S1 followed a straight line over the range of 50 to 90 °C,
for the polymerization temperature of 60 °C when a Kamal model was as shown in Fig. 3. The slope of this plot yielded Ea values of 67 kJ/mol.
This result is quite consistent with the values reported in a previous
work, where the apparent iso-conversional activation energy was de-
scribed [12]. In the bibliography, there are very few reports about the
cyanide polymerization rate, and to our knowledge, this is the first time

Fig. 2. Rate of change of conversion dα/dt vs conversion degree α for aqueous


NH4CN polymerization at different isothermal temperatures. Symbol experi- Fig. 3. Arrhenius plot of the polymerization rate in the first stage vs the tem-
mental values of the conversion rate. Lines: calculated values. perature reciprocal.

4
I. Mas, et al. European Polymer Journal 132 (2020) 109719

reaction.

3.1.3. Theoretical predictions from kinetic experimental parameters


At this point, it is interesting to compare the experimental data with
the simulated predictions calculated from the kinetic parameters gen-
erated from both models. Fig. 4 shows the comparison of experimental
data versus either Kamal or nth-order model calculated values. As
shown in this figure, the deviations are evident, and neither model fits
the conversion data for the two stages of polymerization. For example,
for the highest temperature of 70 °C, an acceptable correlation is found
between the experimental values and the calculated data through both
models, but only at the early stage, where the reaction kinetics are
likely chemically controlled, but later, the theoretical maximum con-
version is higher than that obtained experimentally. This similar ten-
dency is also observed for the reaction at 60 °C, whose experimental
data were also fit to Kamal and nth-order models, as clearly observed in
the plot inset. However, the change in the kinetic model to the nth-
order model does not provide a real improvement to represent the
Fig. 4. Degree of conversion vs reaction time of the aqueous polymerization of process kinetics at a lower temperature of 50 °C.
NH4CN at different isothermal temperatures. Symbol experimental values. On the other hand, from the determination of Arrhenius parameters
Lines: calculated values according to different kinetic models. The inset has the examined in the previous section, it is possible to estimate the con-
same axis legends as that of the full figure.
version rate at another temperature outside the temperature range
considered. Therefore, as indicated in Fig. 3, a value of Rp = 2.6·10−3
that an average value for the activation energy has been calculated for h−1 was calculated for a polymerization temperature of 38 °C. This
this particular case of HCN polymers. In a pioneering work, Ferris and predictable value could be confirmed when compared with further
co-workers indicate that the rate of aqueous cyanide oligomerization to experimental data obtained in previous works from our laboratory
DAMN is not influenced by the presence of nucleophiles and that is only [10,11,16]. Although this set of data at 38 °C is scarce, it is sufficient to
dependent on the initial pH [14]. However, it was recently shown that indicate the known kinetic trend; then, the empirical model based on
ammonium ions influence the cyanide polymerization rate [10], and the Hill function was again applied, as shown in Fig. 4. From this fitting,
Schwartz and Goverde showed that cyanohydrins significantly accel- the following kinetic parameters were obtained: a maximum conversion
erate the rate of HCN oligomerization to DAMN [15]. Therefore, new of 44% and a reaction rate into the chemically controlled stage of
studies are necessary to complete a global vision of cyanide poly- 2.1·10-3 h−1, as indicated in this plot. Both values are expected from the
merization, taking into account the activation energy values to obtain a kinetic analysis done so far.
comprehensive understanding of the real catalysts in this complex

Fig. 5. Evolution of UV–Vis spectra at different reaction times for aqueous NH4CN polymerization at different isothermal temperatures. The UV–Vis spectra of the
NH4CN polymers synthesized at 80 °C were also recorded but for clarity are not shown in the figure.

5
I. Mas, et al. European Polymer Journal 132 (2020) 109719

In general, an evolution in the UV–Vis spectra of the sol fractions


with the polymerization temperature is observed, with those obtained
at 90 °C more complex than those obtained at 50 °C (Fig. 5). Therefore,
the Hill correlation was newly applied to the relationship A230/A300 for
NH4CN polymerization at 50 and 60 °C, and a relatively good fit be-
tween the experimental data and this equation was achieved, with
values of R2 = 0.95 and 0.97, respectively, as illustrated in Fig. 6.
However, the value of the Hill exponent n at 50 °C, n = 0.18 ± 0.10,
cannot be related to that obtained under the same conditions for a
polymerization temperature of 38 °C, n = 0.77 ± 0.12 [10], because it
is lower than what is expected and presents a high error. By contrast,
the value of the Hill exponent found for 60 °C, n = 1.12 ± 0.15, is
indeed in agreement with the value found for 38 °C. However, it is
important to note that this n parameter is different from the reaction
order discussed in the previous section, and it lacks physical meaning.
On the other hand, these absorbance relationships for the fractions
obtained at 70, 80, and 90 °C cannot be correctly fitted to the Hill
function. Very poor R2 values were obtained, R2 ~ 0.5, and the n values
are not consistent between them, and their standard errors are certainly
Fig. 6. Polymerization profiles based on the absorbance ratio for the poly-
great (n = 0.18 ± 0.20, n = 0.29 ± 0.14 and n = 0.15 ± 0.34 for
merization reactions carried out from 50 °C to 90 °C. The lines represent the fit
to the Hill function. The inset has the same axis legends as that of the full figure. 70 °C, 80 °C and 90 °C, respectively), indicating that this approach to
the kinetics of polymerization by means of UV–Vis makes no sense for
these high temperatures. Moreover, the variation in the UV–Vis spectra
3.1.4. Approach to the kinetics by UV–Vis spectroscopy
of the sol fractions during the polymerization process recorded at these
In the previous sections and according to a gravimetric method, the
higher temperatures has significant differences with respect to those
reaction process under study seems to be dominated by different me-
recorded at 50 and 60 °C, and the spectra are notably more complex
chanisms at distinct temperatures, with an autocatalytic reaction
than those described at 38 °C, as explained above.
shifting to an nth-order kinetic process as the polymerization tem-
These results seem to indicate that there is a dissimilarity in the
perature decreases. This dependence should also be reflected in the
pathway of NH4CN polymerization, where the oligomeric intermediates
kinetics of the formation of the soluble oligomers/polymers, i.e., in the
have a different chemical nature dependent on the reaction tempera-
characteristics of the sol fractions obtained in the same way as the gel
ture, and by extension, the formation mechanism and kinetics should be
fractions, taking into account that it has been proposed that the soluble
different, as explained in the previous sections.
NH4CN oligomers/polymers are the intermediate products of the cya-
nide polymerization reactions and that the precipitation of the black
polymers is only a change in the cross-linking grade of the system [10].

Fig. 7. C/N, C/H, C/O and N/O molar ratios for NH4CN polymers synthesized at different temperatures as a function of the conversion. For clarity, the results for
80 °C are not shown in the figure.

6
I. Mas, et al. European Polymer Journal 132 (2020) 109719

3.2. Structural properties of insoluble NH4CN polymers (Fig. 8). The broad band at ~3340 cm−1 can be assigned to hydro-
genated functional groups such as amines (eNH2, eNH), amides (eNH,
Taking into account the results discussed up to this point, it seems amide I) and hydroxyl groups (eOH). At 50 °C, the intensity of the band
evident that the mechanism and kinetics to explain the polymerization at ~1645 cm−1 decreases throughout the polymerization, while the
of NH4CN in water are dependent on the temperature. These differences relative intensity of the band at ~3340 cm−1 remains constant. This
should also be reflected in the structural characteristics of the final same relationship is observed for those samples synthesized at 60 °C
products of the reactions. Thus, a systematic description of the struc- and for those produced at 70 °C, but with a minor range of differ-
tural properties of the NH4CN polymers was made. entiation (Fig. 8). It appears that the evolution of the system at 60 and
70 °C is intermediate between the process that takes place at the lowest
3.2.1. Elemental analysis temperature and the highest temperature considered herein. In con-
Fig. 7 shows the molar ratios C/N, C/H, C/O and N/O as a function trast, at 90 °C, the intensity of the band at ~3340 cm−1 decreases with
of the conversion degree and the reaction temperature for the NH4CN the conversion degree, and the band at ~1645 cm−1 remains unaltered
polymers synthesized here. From the plot of the C/N molar ratio, throughout the polymerization process. Note that all spectra were
Fig. 7a), it is clearly observed that the ratio increases with the con- normalized based on the more intense band for each case.
version degree, indicating denitrogenation processes throughout the Therefore, it is also interesting to calculate the following relation-
reaction progress. This effect is more noticeable at higher reaction ship [I1645/(I3340 + I1645)] × 100 as other and complementary quan-
temperatures when the polymerization is carried out at 90 °C, and the tification of the progress of the NH4CN polymerization, making use of
final product contains less nitrogen than that synthesized at 50 °C, ac- the same equation employed for the EOR. The bottom part of Fig. 9
cording to previous results reported from 75 to 90 °C [12]. Thus, the shows this new relationship, which increases with the conversion de-
high temperatures and conversion degree favour the denitrogenation gree when the NH4CN polymers are synthesized at 90 °C and decreases
processes, and under these conditions, lower limit conversions are when the samples are synthesized at 50 °C. Intermediate values were
achieved. Looking at the C/H molar ratio, Fig. 7b), it is seen at 90 °C calculated for 60 °C and 70 °C, with a decreasing tendency at 60 °C and
that the increase in the conversion degree of the system leads to more a relatively constant or light decrease for the NH4CN polymers syn-
conjugated macrostructures. At the other three temperatures, the effect thesized at 70 °C (Fig. 9). In addition, this relationship might help to
is not as evident, but in general, there is an increase in the conjugation interpret the ratio of conjugation of these macromolecular systems
of the system with the conversion degree, likely through dehy- because the band at ~1645 cm−1 is related to unsaturated double
drogenation reactions. It seems that more conjugated structures are bonds and the band at ~3340 cm−1 is assigned hydrogenated func-
formed at 70 °C, at medium conversion degrees. Finally, from the in- tions, as discussed above. Thus, higher values for this spectroscopic
formation proven by the C/O and N/O molar ratios, it can be seen that parameter may indicate a greater conjugation of the macrostructure
an increase in the temperature favours hydrolytic and/or oxidative considered. In this case, the NH4CN polymers synthesized at 90 °C may
reactions during the polymerization, as was reported previously [12]. present more conjugated macrostructures than those synthesized at
In summary, the 3D plot of Fig. S2 is very enlightening, where it is 50 °C.
observed that the variation in composition undergoes during the In addition, the shape and width of the band centred at
polymerization process is directly dependent on the reaction tempera- ~3340 cm−1 can be related to the presence of –COOH and –CONH2
ture: The higher temperatures enhance the denitrogenation and hy- groups and therefore associated with hydrolysis reactions. For the
drolytic/oxidation process together with dehydrogenation reactions NH4CN polymers synthesized at 50 and 60 °C, it is clearly observed that
throughout the polymerization, leading to macrostructures that are this band is broadened during the polymerization process. This effect is
more oxidized, conjugated and richer in carbon than in nitrogen. also identified for the samples obtained at 70 °C, but it is not so notable.
In contrast, for the polymers produced at 90 °C, the effect is the op-
3.2.2. FT-IR analysis posite; at higher conversion degrees, the intensity and width of this
Fig. 8 shows a representative selection of FT-IR spectra for insoluble band are minor. Taking into account these results and the elemental
NH4CN polymers synthesized at different temperatures and indicates analysis discussed above, the hydrolysis process to generate eCOOH
the conversion degree for each of them. All spectra can be divided into and eCONH2 groups might be mainly during cyanide polymerization at
four main regions: region I (3700–3000 cm−1), region II the lowest temperatures considered here, and other oxidation processes
(2275–2000 cm−1), region III (1825–1000 cm−1) and region IV centred that lead to more conjugated macrostructures might occur at higher
at 660 cm−1. The description and assignation of these bands as well as temperatures.
the evolution of the nitrile groups throughout the polymerization pro-
cess have been largely described previously [2,12,16]. Therefore, we 3.2.3. Solid-state NMR spectroscopy
think that it is not necessary to discuss these aspects again. However, To find other structural differences beyond those commented for
quantitative data for comparative proposes can be obtained from the elemental analysis or FT-IR spectroscopy, the 13C CP MAS-NMR spectra
intensities between the peak centred at ~1645 cm−1, I1645 (assigned to of a selection of NH4CN polymers at limit conversions were recorded, as
conjugated C]C, to (C]N)n structures or to a combination of the two), displayed in Fig. 10. No significant differences were found among the
and the nitrile peak centred at ~2200 cm−1, I2220. Both signals can be three spectra, neither in the recently reported spectra of the samples
used as internal indicators of the extent of reaction (EOR, %) through synthesized at 80 °C [2] nor with those obtained at 38 °C [16]. The
the relationship EOR = [I1645/(I2220 + I1645)] × 100 [12 and internal deconvolution of the resonances of these spectra leads to the accurate
references]. In the upper part of Fig. 9, this relationship is represented interpretation of this complex and heterogeneous system, which was
as a function of the conversion degree. In general, the EOR value in- made in detail in a recent article [2]. This exhaustive analysis indicated
creases with the conversion degree, independent of the temperature; the presence of several functional groups such as carbon in N-con-
however, this parameter is higher for the samples synthesized at higher taining functional groups (> C]NH, extended pyridines, pyrroles,
temperatures. This result is in good agreement with the compositional amides and γ- and δ-lactams), double bonds (C]C) or triple bonds
analysis data discussed above. The degree of hydrolysis of the NH4CN (C^N) and carbon–nitrogen single bonds (CeN). Only slight differ-
polymers and/or the oxygenated composition is higher for the NH4CN ences were found between the intensity of the resonances in the range
polymers synthesized at higher temperatures. of 130–110 ppm. The low intensity in the spectrum of the polymer
Nevertheless, it is noteworthy that the evolution of the bands synthetized at 50 °C could be in agreement with the minor oxidation
centred at ~3340 and ~1645 cm−1 with the conversion degree is op- state of this sample, according to the data obtained by elemental ana-
posite for the NH4CN polymers synthesized at 50 °C and at 90 °C lysis.

7
I. Mas, et al. European Polymer Journal 132 (2020) 109719

Fig. 8. FT-IR spectra of insoluble NH4CN polymers obtained from water polymerization at different temperatures. Note that all spectra are normalized based on the
more intense band, which is dependent on the polymerization temperature and the conversion degree.

Fig. 9. Dependence of both the extension of reaction (EOR) and the relationship
[I1645/(I3340 + I1645)] × 100 calculated from the FT-IR analysis on the con-
version degree of insoluble NH4CN polymers obtained using different tem- Fig. 10. CP MAS solid-state 13C NMR spectra of NH4CN polymers synthesized at
peratures. different temperatures and at limit conversions for each case.

As discussed in previous works, the scarce or null solubility of HCN 3.2.4. XRD powder patterns
polymers in common solvents makes their proper characterization The X-ray diffractograms of a selection of NH4CN polymers were
highly difficult. Thus, alternative techniques should be applied for the recorded (Fig. 11). In Table S3, the main crystallographic parameters
characterization of these solid materials to obtain a better under- are detailed. All samples show a diffraction at ~27.2° and a d-spacing of
standing of their complex nature and the differences between them, ~3.27 Å. These XRD profiles are identical to those described previously
such as X-ray diffraction. for NH4CN polymers synthesized at 80 °C with different conversion
degrees [2]. All these XRD profiles present an extraordinary similarity
with those observed for layered graphitic carbon nitrides g-C3N4, with a

8
I. Mas, et al. European Polymer Journal 132 (2020) 109719

Fig. 11. Representative XRD patterns of NH4CN polymers synthesized at dif-


ferent temperatures and at limit conversions for each case.

peak at 2θ = 27.5° related to the (0 0 2) diffraction typically found in


layered crystal structures [17,18]. These materials have been recently
reviewed due to their singular properties and wide range of applications
[18]. Our results appear to indicate the presence of graphitic-like 2D
structures in the macrostructure of the NH4CN polymers studied here.
In addition, the % of crystallinity of the samples increases when the
polymerization temperature is higher (Table S3), and the peak centre at
27.2° is also narrowed in an identical way (Fig. 11). These last results
might indicate that a higher temperature of polymerization leads to
more ordered NH4CN polymers and therefore to more thermo-
dynamically stable macrostructures.

3.2.5. Study of the insoluble residues after acid hydrolysis of the NH4CN
polymers
It is well known that the acid hydrolysis of insoluble NH4CN poly-
mers generates an important set of polar molecules with interest in
prebiotic chemistry [1,19–21]. However, few studies have been re- Fig. 12. a) Weight of the insoluble residues collected after acid hydrolysis of
ported on the insoluble residues collected after these hydrolysis con- NH4CN polymers synthesized at different temperatures and at limit conver-
ditions. The resistance of the HCN polymers to this treatment may also sions. This percentage was calculated as follows: (mg of insoluble hydrolysed
give an indication regarding the chemical stability of these substances, residue/mg of insoluble NH4CN polymers) × 100. The temperature indicated in
beyond the results obtained by XRD. Fig. 12a) shows the weight per- the x-axis corresponds to the polymerization temperature used for the syntheses
centage of the insoluble residues collected after acid hydrolysis for of the NH4CN polymers. b) and c) Elemental analysis data of the insoluble re-
sidues collected after acid hydrolysis of these same polymers.
samples of NH4CN polymers synthesized at all temperatures under
study for longer reaction times when a limit conversion is achieved. It is
observed that the amount of insoluble hydrolysed residues increased as Fig. 13, where only three regions can be observed). Taking into account
the polymerization temperature increased. This means that NH4CN the FT-IR spectra of the insoluble hydrolysed residues of NH4CN poly-
polymers synthesized at higher temperatures appear to be more stable mers (Fig. 13) and those of the respective unhydrolysed samples
against acid hydrolysis or at least yield a minor amount of soluble or- (Fig. 8), it seems that the main macrostructure remains inalterable, and
ganic materials. only the sensitive functional groups to be hydrolysed are subjected to
In the same figure, compositional data of these insoluble residues the corresponding reactions.
after acid hydrolysis indicate that these samples are more oxidized than In summary, the temperature has a notable effect on the structure of
unhydrolysed samples due basically to a loss of nitrogen, regardless of the insoluble NH4CN polymers. The polymers synthesized at 90 °C
polymerization temperature [Fig. 12b)]. The samples synthesized at present structures that are more oxidised, more conjugated, more
90 °C, although they are initially the most oxidized, are the least oxi- crystalline and ordered and more stable against acid hydrolysis than
dized once hydrolysed [Fig. 12c)]. This result was confirmed by a new those polymers synthesized at lower temperatures.
FT-IR analysis. As expected, the spectra of these hydrolysed residues do
not show the characteristic peak of the NH4CN polymers at 3.3. Some mechanistic, structural and practical aspects about the aqueous
~2200 cm−1 related to nitrile groups, and the band centred at cyanide polymerization
~3340 cm−1 undergoes a shift to 3302 cm−1. Additionally, this band is
clearly broadened due to hydrolysis of nitrile groups to carboxylic acids In this section, a series of clues for the elucidation of the reaction
(please compare the four regions described for the FT-IR spectra of pathways for the production of the insoluble cyanide polymers are
insoluble NH4CN polymers shown in Fig. 8 with those presented in suggested, based on the classic experiments made by Ferris and co-

9
I. Mas, et al. European Polymer Journal 132 (2020) 109719

the chemical transformations that may take place during HCN oligo-
merization, and they found that HCN undergoes a rapid, essentially
irreversible, condensation to DAMN to give oligomeric products [23].
They postulate that redox reactions occur during this process to give
reduced HCN oligomers and diiminosuccinitrile (DISN). These dinitriles
are easily converted to the corresponding hydrolysis products and can
also react among themselves, as has been described in a recent study
based on aqueous DAMN polymerization, where a general picture is
provided; please see Scheme 2 in reference [2]. Therefore, DISN hy-
drolyses to oxalic acid and urea, while DAMN can give the monoamide
of the aminomalononitrile and one equivalent of cyanide or aminoa-
cetonitrile and the monoamide of cyanogen. The monoamide of cya-
nogen may cleave as cyanogen does to yield one equivalent of cyanide;
however, it appears unlikely that a second equivalent of cyanide will be
formed by either reaction pathway, according to Ferris and Edelson
[23]. The monoamide of AMN is known to hydrolyse to the diamide,
and it does not eliminate cyanide, as indicated in Scheme 1. This
Fig. 13. FT-IR spectra of insoluble residues collected after acid hydrolysis of a
selection of insoluble NH4CN polymers synthesized at different temperatures scheme shows that a significant amount of carboxylic acids and amides
and at limit conversions. and other molecular species are formed during HCN polymerization in
water. The production of these carboxylate compounds (carboxylic
acids and amides) by hydrolysis and decyanation are more favourable
workers and other recently reported results. On the other hand, it is
at high temperatures, due to the thermal activation of the system. These
significant that computational studies have indicated the possible ex-
oxidised molecules might lead to other secondary and/or crossing re-
istence of 678 tetrameric structures generated from HCN without con-
actions beyond those shown in Scheme 2 and of the hypothetical
sidering the role of the water [22]. Therefore, taking in account that the
pathways detailed in reference [2], increasing the complexity of the
cyanide polymerization mechanisms seem to be directly dependent on
aqueous cyanide polymerization mechanism. This assumption also
the temperature, such as is indicated above, only some suggestions
might explain the lower conversion obtained for the insoluble NH4CN
about the pathways of reaction are discussed, an only few plots are
polymers at the higher temperatures; and it also might be related on the
shown in an attempt to understand the wide and complex chemical
fact that the UV–Vis spectra of the sol fractions from the higher tem-
space of HCN. Starting from this point, the potential role of cyanide
peratures are more complex than those obtained at the lower tem-
polymers as multifunctional materials and their implications in pre-
peratures. Thus, if the number of possible monomers and oxidised
biotic chemistry are discussed.
molecules which can potentially promote the oligomerization/poly-
Several studies have tried to analyse the catalytic effects on the
merization is increased, the collected amount of the insoluble polymers
condensation reaction of HCN to oligomers. In the first step, HCN
is decreased. This result is due to the increasing of possible reactions in
condenses to form a dimer (iminoacetonitrile), trimer (aminomalono-
the system which might produce a larger diversity of soluble oligomers/
nitrile, AMN) and tetramer (DAMN). One or more of these simple HCN
polymers. In addition, it is important to remember the kinetic study of
derivatives polymerize to yield a complex of substances and may also
Schwartz and Goverde [15], where they demonstrated that different
undergo hydrolytic and/or redox reactions, as indicated in Scheme 1.
cyanohydrin compounds markedly accelerate the rate of HCN oligo-
At this point, if it is difficult to know with certainty the key monomer
merization, both in homogeneous solution as well as in the frozen
that originates the growing chains, it can be truly more problematic to
aqueous state. All these results suggest that hydroxylated compounds,
elucidate what compound acts as the catalyst during this reaction.
generated from hydrolysis and/or oxidation reactions, can have a cat-
However, all the studies coincide in pointing to AMN or DAMN as the
alytic effect during the polymerization process. At the higher tem-
true monomers that give rise to the growth of the polymer chains.
peratures considered here, the hydrolysis and/or oxidation reactions
Therefore, as an example in Scheme 2, a general picture of the poly-
seem highly probable considering the structural data reported in the
merization reaction in aqueous medium is presented. Here, it is possible
previous sections, and the discussion indicated above. Since the auto-
to distinguish two polymerization routes. On the one hand, one route is
catalytic behaviour must be considered at the higher temperatures
based on the formation of a linear polyimine through the decomposi-
studied, some of the oxidised molecules specified above must be the key
tion step of the trimer AMN to give a carbene with the elimination of
in the autocatalytic process. However, due to the complexity of the
hydrogen cyanide. It seems more likely that this polyimine, formulated
system and the current information (computational as well as analy-
here with an oligomeric formula of C12N12H12, can yield the different
tical/spectroscopic), it is very difficult to indicate which of them can be
heterocyclic macrostructures indicated in this scheme according to the
the kingpin in the autocatalytic behaviour at the higher temperatures.
mechanism explained in detail in a recent work [2]. In addition, in this
Moreover, other possibilities also exist; for example, polymeric pro-
study, the hydrolysis of polyamine and subsequent intramolecular cy-
ducts formed during the precipitating polymerization could act as cat-
clization are postulated to give a new family of oxygenated heterocycles
alysts. In this sense, the autocatalytic character of HCN polymers is
onto the polymeric chains. On the other hand, Thissen and co-workers
known, and it has been demonstrated experimentally by seeding a to-
have proposed a new mechanism for AMN polymerization from aqu-
luene solution of HCN with a dark HCN polymer and finding that it
eous solutions, giving in this case polymeric films, as indicated in this
promotes the formation of a new polymer [internal references in 12].
scheme [8]. They suggest the formation of a linear polyamine,
At this point, the results presented herein offer, on the one hand,
C12N12H12, by a polyaddition between amine and cyano groups, which
that high temperatures lead to highly conjugated and oxidised and re-
evolves to a polyaminoimidazole; they also propose its hydrolysis to
latively largely ordered HCN-derived polymeric structures. It is also
give a new macrostructure with carbonyl groups in the imidazole rings,
likely that these high temperatures might lead to changes in the mor-
C12N9H3O3, according to experimental compositional data. Logically,
phology, size and density of the final cyanide polymerization products
the hydrolysis reaction of polyamine must be considered, as well as
(preliminary results of further work in progress are in agreement with
other hypothetical macrostructures such as those presented in this
this assumption). And on the other hand, the temperature has a sig-
scheme.
nificant influence in the kinetics of the production of HCN-derived
Returning to Scheme 1, Ferris and Edelson scrutinized in high detail
polymers and subsequently in the mechanistic pathways. The great

10
I. Mas, et al. European Polymer Journal 132 (2020) 109719

Scheme 1. Oligomerization of cyanide and hydrolysis reactions according to pioneering studies by Ferris and co-workers [14,22].

complexity of the cyanide polymerization pathways in aqueous en- The confusion is still greater when the nature of the precursors that lead
vironment is well-known, but here we demonstrate and for the first to this autocatalytic behaviour is unknown. However, in spite of these
time, that this complexity is greater than expected, since the mechan- newly revealed open questions indicated here, some notable practical
isms of HCN-derived polymer production are temperature dependent. suggestions can be made based on the new structural data obtained: i)

Scheme 2. Hypothetical pathways for the formation of HCN polymer from AMN according to recent works in the literature [2,8].

11
I. Mas, et al. European Polymer Journal 132 (2020) 109719

the versatility of the HCN-derived polymers synthesis allows one to prebiotic chemistry and astrobiology.
tune the properties of the polymers by a correct selection of the reaction
conditions, such as temperature and time; ii) recently this polymeric CRediT authorship contribution statement
system has been compared to graphitic carbonitrides due to the partial
presence of graphite-like 2D structures in NH4CN and DAMN polymers Itziar Mas: Validation, Formal analysis, Investigation. José L. de la
[2]. The high temperature favours the arrangement in graphite-like 2D Fuente: Conceptualization, Methodology, Formal analysis, Writing -
structures, such as was shown by XRD. This result, and the more con- original draft, Writing - review & editing. Marta Ruiz-Bermejo:
jugated structures of the polymers obtained at the higher temperatures, Conceptualization, Methodology, Validation, Formal analysis,
encourage us to study the potential applications of these polymers as Resources, Writing - original draft, Writing - review & editing,
new multifunctional materials in analogy to the well-known g-C3N4 Supervision, Funding acquisition.
graphitic carbonitrides [18]; iii) the fine-tuning of the textural prop-
erties by changes in the morphology, size or density, prompted by the Declaration of Competing Interest
temperature conditions, can lead to the obtention of HCN-derived
polymeric nanoparticles which can be incorporated in other polymeric The author(s) declare no competing interests.
matrices which could have a relevant impact in the field of polymer
composites, as was previously proposed [2]; iv) the thermal conditions Acknowledgements
might also be able to tune the properties of the HCN-derived polymeric
coatings, currently promising materials with potential biomedical ap- The authors used the research facilities of the Centro de
plications [4,5]; v) the HCN is a ubiquitous molecule in the Universe Astrobiología (CAB) and were supported by the Instituto Nacional de
and it is considered as a preferential prebiotic reagent since the HCN Técnica Aeroespacial “Esteban Terradas” (INTA), by the project
polymers are proposed as main prebiotic precursors of important bio- ESP2017-89053-C2-2-P from the Spanish Ministerio de Ciencia,
molecules (see reference [1] and the references therein). In addition, Innovación y Universidades and by the Spanish State Research Agency
the HCN polymers are considered the oldest organic materials in the (AEI) project MDM-2017-0737 Centro de Astrobiología (CSIC-INTA),
Solar System. Now, it is revealed that they provide potential properties Unidad de Excelencia María de Maeztu. Additionally, the authors are
as semiconductors and catalysts and subsequently they will likely have grateful to Mª Teresa Fernández for performing the FT-IR spectra and
an important role in the development of new reactions. Thus, the HCN- XRD measurements and to David Hochberg for the detailed revision of
derived polymers may lead to the discovery of new unexplored horizons the manuscript.
in prebiotic chemistry, beyond the traditional ways marked by Oró's
synthesis of adenine [24]. Data availability

4. Conclusions The raw/processed data required to reproduce these findings cannot


be shared at this time due to technical or time limitations.
It has been demonstrated that the temperature has a significant
effect on NH4CN polymerization. The fit to phenomenological kinetic Appendix A. Supplementary material
models, autocatalytic or nth order, is conditioned by the reaction
temperature. Moreover, this result is reflected in good agreement Supplementary data to this article can be found online at https://
through the structural features and properties of the insoluble NH4CN doi.org/10.1016/j.eurpolymj.2020.109719.
polymers synthesized. Therefore, an autocatalytic mechanism seems to
prevail at higher temperatures, and the respective insoluble polymers References
present highly oxidized, conjugated and ordered macrostructures;
therefore, they are more thermodynamically stable. In contrast, an nth- [1] M. Ruiz-Bermejo, M.-P. Zorzano, S. Osuna-Esteban, Simple organics and biomo-
order kinetic model is followed for these precipitation polymerizations nomers identified in HCN polymers: An overview, Life 3 (2013) 421–448.
when they react at medium temperatures, and in this case, the corre- [2] M. Ruiz-Bermejo, J.L. de la Fuente, J. Carretero-González, L. García-Fernández, M.
R. Aguilar, A comparative study of HCN polymers synthesized from NH4CN or
sponding polymeric materials are more labile and disordered, according DAMN polymerization in aqueous media: New perspectives for prebiotic chemistry
to both XRD and chemical stability measurements. and material science, Chem. Eur. J. 25 (2019)11437–11455.
For the first time, a reliable value for the activation energy of cya- [3] X. Zhou, Y. Fang, Y. Su, C. Ge, B. Jin, Z. Li, S. Wu, Preparation and characterization
of poly-hydrogen cyanide nanofibers with high visible light photocatalytic activity,
nide polymerization, between 50 and 90 °C, has been obtained from the Catal. Commun. 46 (2014) 197–200.
initial polymerization rate, where the polymerization is chemically [4] H. Thissen, A. Koegler, M. Salwiczek, C.D. Easton, Y. Qu, T. Lithgow, R.A. Evans,
controlled. The empirical model proposed herein allows an excellent Prebiotic-chemistry inspired polymer coatings for biomedical and material science
application, NPG Asia Mater. 7 (2015) e225, https://doi.org/10.1038/am.2015.
description of the two regions, chemical- and diffusion-controlled, that 122.
characterize the kinetic behaviour of this extraordinarily complex [5] D.J. Menzies, A. Ang, H. Thissen, R.A. Evans, Adhesive prebiotic chemistry inspired
polymerization reaction, which presents diffusion limitations. This ap- coatings for bone contacting applications, ACS Biomater. Sci. Eng. 3 (5) (2017)
793–806.
proach based on the Hill equation is simple, easy to use and in very
[6] V. Ball, R.J. Tohc, H. Nicolas, N.H. Voelcker, H. Thissen, R.A. Evans,
good agreement with the experimental data, and it could be applied to Electrochemical deposition of aminomalonitrile based films, Colloids and Surfaces
other experimental polymerization conditions. Moreover, the descrip- A 552 (2018) 124–129.
tion of the kinetically and chemically controlled stages and the limiting [7] J. Junga, D.J. Menzies, H. Thissen, C.D. Easton, R.A. Evans, R. Henry, A. Deletic, D.
T. McCarthya, New prebiotic chemistry inspired filter media for stormwater/grey-
diffusion conversions have been accurately described from 38 to 90 °C. water disinfection, J. Hazard. Mater. 378 (2019) 120749.
From an industrial standpoint for process modelling and optimiza- [8] R.J. Toh, R. Evans, H. Thissen, N.H. Voelcker, M. d’Ischia, V. Ball, Deposition of
tion, this methodology is usually preferred. However, it is important to aminomalononitrile-based films: kinetics, chemistry, and morphology, Langmuir 35
(30) (2019) 9896–9903.
note that the reactions carried out herein are not stirred to favour [9] M.R. Marín-Yaseli, C. Cid, A.I. Yagüe, M. Ruiz-Bermejo, Detection of macro-
precipitation processes. Thus, new experiments must be performed to molecular fractions in HCN polymers using electrophoretic and ultrafiltration
elucidate the full role of the diffusion stage when NH4CN polymeriza- techniques, Chem. Biodiver. 14 (2017) e1600241, https://doi.org/10.1002/cbdv.
201600241.
tions are driven by a stirring method. All the results shown, both the [10] M.R. Marín-Yaseli, M. Moreno, J.L. de la Fuente, C. Briones, M. Ruiz-Bermejo,
kinetic and the structural and spectroscopic data, may be very useful in Experimental conditions affecting the kinetics of aqueous HCN polymerization as
designing new materials dependent on the desired applications and in revealed by UV-vis spectroscopy, Spectrochim. Acta Mol. Biomol. Spectrosc. 191
(2018) 389–397.
beginning to understand the real role of cyanide polymerization in

12
I. Mas, et al. European Polymer Journal 132 (2020) 109719

[11] M. Ruiz-Bermejo, J.L. de la Fuente, M.R. Marín-Yaseli, The influence of reactions carbonitrides from biotic molecules for photoredox transformations, Angew. Chem.
conditions in aqueous HCN polymerization on the polymer thermal degradation Int. Ed. 56 (2017) 6627–6631.
properties, J. Anal. App. Pyrolysis 124 (2017) 103–112. [18] M. Inagaki, T. Tsumura, T. Kinumoto, M. Toyoda, Graphitic carbon nitrides (g-
[12] A. Fernández, M. Ruiz-Bermejo, J.L. de la Fuente, Modelling the kinetic and C3N4) with comparative discussion to carbon materials, Carbon 141 (2019)
structural properties evolution of a versatile reaction: the aqueous HCN poly- 580–607.
merization, Phys. Chem. Chem. Phys. 20 (2018) 17353–17366 and references [19] M.R. Marín-Yaseli, C. Mompéan, M. Ruiz-Bermejo, A prebiotic synthesis of pterins,
herein. Chem. Eur. J. 21 (2015) 13531–13534.
[13] D.S. Achillas, A review of modeling of diffusion controlled polymerization reac- [20] M.R. Marín-Yaseli, E. González-Toril, C. Mompeán, M. Ruiz-Bermejo, The role of
tions, Macrom. Rapid Commun. 25 (2007) 319–347. the aqueos aerosols in the “Glyoxylate Scenario”: An experimental approach, Chem.
[14] J.P. Ferris, D.B. Donner, A.P. Lobo, Possible role of hydrogen cyanide in chemical Eur. J. 22 (2016) 12785–12799.
evolution: The oligomerization and condesantion of hydrogen cyanide, J. Mol. Biol. [21] E. Borquez, H.J. Cleaves, A. Lazcano, S.L. Miller, An investigation of prebiotic
74 (1973) 511–518. purine synthesis from the hydrolysis of HCN polymers, Origin Life Evol. Biosph. 35
[15] A.W. Schwartz, M. Goverde, Acceleration of HCN oligomerization by formaldehyde (2005) 79–90.
and related compounds: Implications for prebiotic synthesis, J. Mol. Evol. 18 (1982) [22] S. Nandi, D. Bhattacharyya, A. Anoop, Prebiotic chemistry of HCN tetramerization
351–353. by automated reaction search, Chem. Eur. J. 24 (2018) 4885–4894.
[16] M. Ruiz-Bermejo, J.L. de la Fuente, C. Rogero, C. Menor- Salván, S. Osuna-Esteban, [23] J.P. Ferris, E.H. Edelson, Mechanism of condensation of cyanide to HCN oligomers,
J.A. Martín-Gago, New insights into the characterization of insoluble black HCN J. Org. Chem. 43 (1978) 3989–3995.
polymers, Chem. Biodiver. 9 (2012) 25–40. [24] J. Oró, P. Kimball, Synthesis of purines under possible primitive Earth conditions. I.
[17] C. Yang, B. Wang, L. Zhang, L. Yin, Wang X. German, Synthesis of layered Adenine from hydrogen cyanide, Arch. Biochem. Biophys. 94 (1962) 217–227.

13

You might also like