1. Introduction
Biosensors typically employ the natural enzyme HRP to specifically detect target molecules. The enzyme’s catalytic POD activity, combined with a signal generator, converts the biochemical interaction into a measurable signal [
1,
2]. Nevertheless, natural enzymes present several limitations as a catalyst including low thermal and environmental stability, sensitivity to environmental conditions, non-conductivity, and the difficulty and expense of synthesis and purification, which has led to the utilisation of nanozymes [
3,
4]. In this regard, various types of artificial enzymes have been introduced including organic materials (for instance, fluorescein/quinone derivatives), metal organic framework-based materials (like Cu
2+–bipyridine complexes), metal-based materials (such as Au, Pt, and Mo), and particularly Fe-based materials [
5,
6,
7,
8].
PB, (Fe(III)
4[Fe(II)(CN)
6]
3), is one of the oldest discovered coordination polymer materials, consisting of ferric and ferrocyanide ions with a face centred cubic unit cell, where the nitrogen coordinates to Fe
3+ and carbon coordinates to Fe
2+. The physicochemical properties of PB including sorption, photo-magnetism, host–guest interactions, redox behaviour, and intrinsic porosity have been observed to result in a wide range of reported applications. It is noteworthy that PB materials have been approved by the U.S. Food and Drug Administration (FDA) for use as an antidote for humans in the event of poisoning by Cs
+ and Tl
+. Furthermore, PBNPs are employed in biomedical applications as an effective photothermal agent for photothermal therapy in cancer cells. They have also been utilised in gas capture and storage, Li
+/Na
+ batteries, sensors, theragnostic, and in nanoprobes for magnetic resonance imaging [
9,
10,
11].
PB has been demonstrated to possess excellent optical and electrochemical properties, which have led to its exploitation in a variety of catalytic-based applications [
12]. In this context, the Fe core is capable of acting as a catalytic centre due to its ability to electrochemically switch between multiple oxidation states including the fully oxidised, partially oxidised, and reduced forms of PB, which are Prussian yellow, Berlin green, and Prussian white, respectively [
13,
14]. Due to their strong electron transport ability, PBNPs are capable of mimicking the multienzyme-like activity exhibited by POD, catalase (CAT), and superoxide dismutase (SOD), indicating it as a promising enzymatic mimetic inorganic material [
15].
In biosensor applications that detect analytes by linking them to enzymes via biomolecules (such as antibodies), thus generating a chromatic signal, the interaction between the nanozyme and these biomolecules is of great importance [
16,
17]. Therefore, to enhance biomolecule binding, nanozymes can be decorated with aminated polymers, such as PEI, which is prompting a growing need to investigate the POD-mimicking activity of these modified nanoconstructs. In this regard, Liang et al. synthesised PEI coated cubic shape PBNPs with a 50 nm edge length by the one-pot thermal reduction protocol [
18]. These particles were reported as a pH switchable nanozyme displaying alternatively POD-like and CAT-like catalytic properties while shifting the pH from acidic to basic, respectively. Interestingly, the highest POD-like activity was observed at pH 5. Using a different approach, Pandey et al. synthesised PEI mediated copper-iron hexacyanoferrate (Cu/Fe HCFs) and nickel-iron hexacyanoferrate (Ni/Fe HCFs) nanoparticles by reducing potassium ferricyanide by PEI at 60 °C [
19]. They optimised the metal ratios, finding that a 1:1 Cu:Fe molar ratio and a 1:5 Ni:Fe molar ratio resulted in nanoparticles with electrochemical properties similar to PB. Additionally, both Cu/Fe HCFs and Ni/Fe HCFs displayed POD-like activity, with Cu/Fe HCFs exhibiting a higher catalytic performance compared to Ni/Fe HCFs. Furthermore, only a few studies have described the use of electrodes modified with PB and PEI layers. Pchelintsev et al. reported PEI/PB coated screen printed electrodes, subsequently modified with glucose oxidase (GOx), as an amperometric biosensor for glucose [
20]. Pajor-Świerzy et al. employed PEI to anchor PB/poly(allylamine hydrochloride) (PAH) multilayers on gold electrodes, thereby enhancing the electrochemical properties of the film through the addition of a conductive polymer, poly(3,4-ethylenedioxythiophene)–poly(styrenesulphonate) (PEDOT:PSS). In this case, the incorporation of the conductive polymer into the multilayers resulted in an improvement in the electrocatalytic sensitivity of hydrogen peroxide (H
2O
2) detection [
21]. Additionally, PB/PEI–graphene multilayer films were reported to exhibit excellent electrochromic properties and good electrocatalysis towards H
2O
2 [
22].
The majority of the PB/PEI NPs employed in nanozymatic studies have been synthesised through either the thermal reduction of ferricyanides in the presence of ferric salts and PEI, or with co-precipitation techniques in the presence of PEI (in situ/ex situ), with ferrocyanide/ferric salt injection. Nevertheless, little is known about the influence of the fundamental characteristics of PB/PEI NPs, which are derived from the synthesis technique, on the nanozymatic-POD-like activity. In this study, three distinct protocols were employed to synthesise a variety of PB/PEI NPs. The first protocol involved a controlled co-precipitation approach, whereby the controlled in situ synthesis of PBNPs was conducted within a PEI bath. The second protocol entailed a thermal reduction method, wherein the reduction of ferricyanide was achieved in the presence of PEI and ferric salt. Finally, the third protocol involved a vortex synthesis approach. For the latter, the PB/PEI NPs were precipitated within a custom-built vortex reactor. To the best of our knowledge, this is the first time that PB/PEI NPs synthesised by protocols (1) and (3) have been reported. The particles were characterised by transmission electron microscopy (TEM), X-ray diffraction (XRD), ultraviolet–visible spectroscopy (UV–Vis), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectroscopy (FTIR), scanning electron microscopy (SEM) coupled with energy dispersive X-ray spectroscopy (EDX), and cyclic voltammetry (CV), and their nanozymatic-POD-like activity was investigated in order to elucidate the influence of the physical, chemical, and electrochemical properties of the particles, which were derived from the synthesis protocol, on their POD-like activity.
3. Results and Discussion
A controlled coprecipitation technique was employed to synthesise the bare PBNPs, designated as
Cb. In this technique, it is crucial to control the precursor concentration, injection rate, and the reaction medium temperature in order to obtain the desired particle size and shape. The manipulation of the aforementioned parameters, either collectively or individually, results in alterations to the shape, size, and Fe/Na. For instance, an increase in the precursor injection rates, while maintaining the other conditions constant, leads to a reduction in particle size and the generation of a broad range of particle sizes. Accordingly, these parameters were precisely maintained in accordance with the specifications outlined in
Section 2.1. The
Cb nanoparticles exhibited a cubic shape and a particle size of 99 ± 13 nm (edge/side length of the cubes), as illustrated in the TEM image (
Figure 1a). The structural formula of
Cb was determined to be Na
0.32Fe[Fe(CN)
6]
0.83, calculated by the Fe/Na ratio in the EDX analysis (
Table S1). A modification of this co-precipitation technique, which was used to synthesise
Cb, was employed to synthesise in situ PB/PEI NPs (
C1) by nucleating and growing the nanoparticles within an acidic (pH 1) PEI solution. The negatively charged Na
4Fe(CN)
6 is capable of attracting the positively charged PEI chains, facilitating the nucleation of PB/PEI NPs in the presence of Fe
3+ ions. However, particle nucleation and growth were altered by the presence of the polymer in the particle synthesising medium. Consequently, the
C1 PB/PEI NPs exhibited a formula of Na
0.23Fe[Fe(CN)
6]
0.81 and a cubic structure with an uneven edge surface and a broader range of particle sizes (92 ± 23 nm, as shown in
Figure 1b). A custom-made vortex reactor (
Scheme S1) was employed to synthesise the
V1 PB/PEI NPs. To the best of our knowledge, this is the first instance of this protocol being used to synthesise Prussian blue-type nanoparticles. Nevertheless, this technique has already been applied to a number of distinct nanoparticle types with prominent examples including co-precipitated poly(vinyl pyrrolidine)-coated superparamagnetic magnetite nanoparticles [
25]. In this process of
V1 synthesis, droplets of the precursors, PEI/Na
4Fe(CN)
6 and FeCl
3, were subjected to vortexing inside the reactor, which was rotated at 1600 rpm. The shear stress applied to the precursor solution led to the formation of smaller ill-defined nanoparticles (19 ± 4 nm;
Figure 1c). Furthermore, the EDX analysis revealed that the nanoparticles exhibited the lowest alkali metal percentage, with a formula of Na
0.11Fe[Fe(CN)
6]
0.78.
Three types of particles with different shapes and sizes were synthesised by the thermal reduction of [Fe(CN)
6]
3− in the presence of PEI, which is attracted electrostatically to [Fe(CN)
6]
3−. The [Fe(CN)
6]
3− is then reduced to [Fe(CN)
6]
4−, followed by the formation of PB/PEI NPs after complexation with Fe
3+. PB/PEI NPs
T1 with formula K
0.15Fe[Fe(CN)
6]
0.78 were synthesised with high precursor concentrations (see
Section 2.2. for details), resulting in the generation of highly aggregated non-uniform cubic/spherical particles with sizes of 26 ± 4 nm (
Figure 1d). By using lower precursor concentrations and increasing the ionic strength by adding KCl to the PB/PEI NPs, we obtained
T2 particles, K
0.32Fe[Fe(CN)
6]
0.83, which were cubic-shaped and 60 ± 11 nm in size (
Figure 1e). The room temperature reduction of [Fe(CN)
6]
3− into [Fe(CN)
6]
4− and reaction without refluxing over 60 h produced PB/PEI NPs
T3 (
Figure 1f), which displayed a cubic shape, 49 ± 10 nm size, and the formula of K
0.14Fe[Fe(CN)
6]
0.79. In the case of
T3, a slower kinetic rate synthesis process was trialled to ascertain the viability of synthesising PB/PEI without refluxing. However, under the aforementioned conditions in
Section 2.2, it took nearly 60 h to generate a PB crystalline phase, which is considerably longer than the refluxing approach, which only takes 30 min. It is worth noting that the synthesis of the nanoparticles by this thermal reduction protocol is highly sensitive to the precursor saturation/volume, pH, and polymer percentage. Thus, for these three types of nanoparticles, particularly for
T1, the solute concentration may exceed the critical nucleation point when subjected to high solute concentrations. This can result in the rapid formation of a high percentage of small nuclei through explosive nucleation. In general, with regard to the considered NP types, the particles synthesised with low kinetic rates for a long-time reaction process tended to result in higher particle sizes. However, under conditions of continuous precautionary feeding, the NPs could be grown in a controlled manner compared to short-time, faster kinetic rate synthesis methods.
Of the three types of particle synthesis protocols, the vortex protocol shows considerable promise for scaling up through the use of continuous precursor injection and continuous product withdrawal from the vortex reactor. The co-precipitation technique presents a significant challenge when attempting to scale up, as the parameters must be optimised for different batch sizes. However, the sensitivity of the particle size and shape, coupled with the necessity for an external energy input to drive the temperature evolution, makes thermally-synthesised particles particularly unfavourable for large-scale production.
The XRD matching pattern for all six types of particles, as seen in
Figure 2a,b, can be attributed to the standard PB crystal JCPDS card 73-0689 with a cubic crystal system with an Fm-3 m space group, and the main peaks around 2θ of 17.452°, 24.778°, 35.326°, 39.660°, 43.630°, 50.820°, 57.337°, 66.322°, and 69.172° [
26,
27]. The crystallinity of the PEI-mediated in situ precipitated
C1 nanoparticles remained roughly unchanged compared to the bare
Cb NPs (
Figure 2a). The diffraction patterns were refined using the pattern matching method to obtain accurate information of the crystalline materials (see
Table S2). The lattice parameter-a of the
C1 NPs (10.16625 Å) was smaller than that of the bare PBNPs (10.17163 Å). This reduction can be attributed to the external pressure exerted by the PEI polymer during the nucleation and growth of the NPs, which compresses the crystal lattice [
28,
29]. The electrostatic interaction and Fe-N coordination between the crystal and PEI can lead to a compressive contraction of the lattice planes, further contributing to the overall lattice compression. It is noteworthy that the mean microstrain of these two particles exhibited nearly identical values of 0.09% and 0.08%, indicating a minimal prevalence of crystal defects in their structure. This observation lends support to the view that the in situ synthesis of PBNPs does not result in any substantial crystal defects or microstrains. On the other hand, the
V1 NPs exhibited the highest lattice parameter (10,17931 Å), together with the highest microstrain of 0.90% among all particles. The reduced particle size of
V1 resulted in the constrained arrangement of a high number of atoms at the surface, leading to elevated surface energies. This effect contributed to an increase in crystal defects and a broad distribution of microstrains [
30]. Moreover, the rapid nucleation of nanoparticles in this vortex protocol gives rise to enhanced crystal distortion, characterised by elevated microstrains, and a notable reduction in crystallinity (
Figure 2a).
The XRD patterns of the particles that were synthesised by thermal reduction are shown in
Figure 2b. The microstrain value and the lattice parameter of these nanoparticles were 0.66%/10.15566 Å, 0.18%/10.16788 Å, and 0.08%/10.15285 Å for
T1,
T2, and
T3, respectively. The same phenomenon of high microstrains in smaller particle sizes was demonstrated for the
T1 nanoparticles, which exhibited 0.66% of microstrain. Furthermore, the combination of temperature with high reaction kinetic rates was a contributing factor to the high microstrain observed in both
T1 and
T2 compared to the microstrain values of the
T3 nanoparticles (0.08%), which were synthesised at room temperature with slower kinetic rates. The crystallite sizes of the particles were estimated by pattern matching to be 91, 62, 23, 18, 73, and 45 nm for
Cb,
C1,
V1,
T1,
T2, and
T3, respectively (see
Table S2).
As illustrated in
Figure 3a, the conventional FTIR vibrational peaks of the bare PBNPs
Cb were attributed to the bands at 3210/3646, 3596, 1604, and 1421 cm
−1 for the OH stretching vibrations in the absorbed H
2O (bonded OH/free OH), OH stretching vibrations in the crystallised H
2O, the OH bending vibrations in the absorbed H
2O, and OH bending vibrations in the crystallised H
2O, respectively [
31]. An intense band for the CN stretching vibration could be observed at 2063 cm
−1, accompanied by the Fe
2+-CN stretching and the bending vibrational bands at 598 and 498 cm
−1, respectively [
10,
32]. All of the aforementioned vibrational bands were present in all of the PB/PEI NPs (
Figure 3b and
Table S3), providing evidence for the synthesis of PB. In addition, a number of peaks were characteristic of only the PB/PEI NPs. A vibrational band appeared at 3249 cm
−1 in both
C1 and
T1 and at 3216 cm
−1 in
V1 (
Figure 3c). This peak can be attributed to the N-H stretching vibrational band that appeared due to the presence of PEI. However, this band might overlap with that of the OH stretching band, which made its identification challenging in
T2 and
T3. A weak band around 1650 cm
−1 in all of the PB/PEI particles could be assigned to the bending vibrations of N-H bonding (
Figure 3d) [
32,
33]. This band was red-shifted compared to the pure PEI polymer and could be clearly identified in the
C1 and
V1 particles at 1650 cm
−1 and 1654 cm
−1, respectively. Furthermore,
C1 and
V1 exhibited the N-H wagging-bending vibrational peak at 783 and 778 cm
−1, respectively (
Figure 3e). The C-H bending vibrational band was also observed at approximately 1375 cm
−1 for all of the PB/PEI NPs, and another C-H bending vibrational band could be seen in
Cb,
V1, and
T2 at 1476, 1475, and 1470 cm
−1, respectively (
Figure 3f) [
34,
35]. The aforementioned data indicate that the presence of PEI in the particles was more pronounced for the in situ coprecipitated and vortex particles than in the particles produced through the thermal process.
XPS characterisation of the bare PBNPs
Cb revealed the typical binding energies (
Figure 4a). The most prominent XPS peaks at 55.1 eV, 93.1 eV, 285.0 eV, 398.1 eV, 532.4 eV, and 708.9 eV, and 1073.2 eV were associated with the binding energies of the electrons of Fe 3p, Fe 3 s, C 1 s, N 1 s, O 1 s, Fe 2p, and Na 1 s, respectively [
36]. Similar XPS peaks were obtained for
C1 and
V1. However, for the
T1,
T2,
T3 particles, a K 2p peak around 294.2 eV was observed instead of the Na 1 s peak at 1073.2 eV (
Figure S1). High-resolution XPS analysis was conducted on the Fe 2p electrons in the particles in order to obtain the valence information of Fe, which is the main component involved in the POD-like activity in PB. The XPS spectra of
Cb particles were deconvoluted into five distinguishable peaks, as illustrated in
Figure 4b. The three principal peaks, situated at 708.9 eV, were indicative of the Fe
2+ 2p
3/2 valence state, while the peaks at 710.2 eV and 712.5 eV were indicative of the Fe
3+ 2p
3/2 valence states, respectively. Two additional satellite peaks were identified at 715.1 eV and 720.0 eV [
37,
38]. The XPS spectra of the five types of PB/PEI NPs exhibited a similar pattern, with five deconvolution peaks situated at approximately the aforementioned positions. The valence information obtained from XPS revealed that the Fe
2+/Fe
3+ ratio was 0.94, 0.88, 1.37, 1.31, 1.28, and 1.25, calculated according to the above-mentioned three peaks in
Cb,
C1,
V1,
T1,
T2, and
T3, respectively. The synthesis protocols, which included controlled co-precipitation, thermal reduction, and the vortex method, demonstrated a low to high Fe
2+/Fe
3+ ratio trend.
High resolution narrow scan XPS analysis was performed on the N 1 s electrons in the particles to determine the presence of PEI on their surface. The two main deconvolution peaks at 398.1 and 399.4 eV in the
Cb particles, as shown in
Figure 5a, corresponded to the C≡N bond coordinated to Fe
2+ and Fe
3+ in the PB structure, respectively [
18,
39,
40]. These two peaks also appeared in all types of PB/PEI NPs, but shifted ±1 eV (
Figure 5b–f). The minor deconvolution peak in
Cb at 402.5 eV may be due to the oxidisation of the N atom [
40]. Nevertheless, in comparison to the three deconvolution peaks (A, B, C) observed in the N 1 s XPS spectra of
Cb, four deconvolution peaks (including an additional D) were present in
C1,
T1,
T2, and
T3, while five deconvolution peaks (additional D and E) appeared in
V1, attributed to the incorporation of PEI. The peak D at 400.8 eV in
C1,
T1,
T2, and
T3 and at 401.4 eV in
V1 can be attributed to the NH from the PEI. Additionally, the peak at 400.1 eV in
V1 can be attributed to the amine N-R/N-R2 groups in the PEI [
23,
41,
42,
43]. Furthermore, the protonated amine peaks were also anticipated to be present at approximately 402.1 eV, which may overlap with the peak at 402.5 eV [
18]. Nevertheless, the increase in the peak area at 402.5 eV indicates the additional N (from PEI) in all PB/PEI NPs, confirming the presence of PEI. It can be reasonably inferred that the increase in N concentration attributable to PEI was approximately 5.3, 6.5, 4.0, 3.7, and 25.3% in
C1,
T1,
T2,
T3, and
V1, respectively, compared to
Cb.
The electrochemical behaviour of the PBNPs and PB/PEI NPs was studied by cyclic voltammetry by drying 10 µL of the nanoparticles (0.1 mg/mL in water) on the working electrode of an SPCE and using 0.1 M KCl (pH 4) as the electrolyte. The conventional electrochemical switching of oxidation (partially/fully) and reduction of PB into Berlin green (BG)/Prussian yellow (PY) and Prussian white (PW), respectively, can be explained by the following chemical equations [
14,
15,
44]. In these equations, A represents Na or K, which originate from the synthesis, and B represents K, which originates from the electrolyte.
The conventional electrochemical behaviour of PB was clearly evident in all nanoparticles, which showed two pairs of quasi-reversible peaks at potentials between 0 and 0.23 V for PB to PW (redox couple 1) and 0.71 V and 0.86 V for PB to BG (redox couple 2; BG—{Fe(III)
3[Fe(III)(CN)
6]
2[Fe(II)(CN)
6]}
−;
Figure 6) [
15]. When comparing
Cb and
C1, which were synthesised by co-precipitation,
C1 displayed larger ΔE and higher peak currents for the first redox couple, indicating lower reversibility but enhanced electron transfer (
Table S4). This was attributed to the protonated amine groups in the polymer, which acted as a bridge to transfer electrons and supported the increase in current, but at the same time affected the PB structure. Moreover, the elevated symmetry of the peak shapes representing the first redox couple in Cb suggests that the electron transfer processes in Cb are more reversible than in C1. The
V1 particles were also nucleated by precipitation, but through the vortex mixing protocol. The resulting particles were smaller than
Cb and
C1 and displayed smaller ΔE and higher peaks than
Cb for the first redox pair. In contrast, the redox potentials observed in the second redox couple were nearly identical for
Cb,
C1, and
V1 (
Figure 6a), reaching the perfect reversibility region below 59 mV (for a one-electron transfer), but the peaks were wider and lower for
V1. This was probably due to the high polymer loading in
V1 (see XPS data in
Figure 5c).
Figure 6b depicts the CVs of PB/PEI NPs produced by the thermal reductive protocol, which displayed two redox waves similar to those observed for the co-precipitated bare particles
Cb. It can be generally expected that, for a fixed mass of NPs, smaller particles will exhibit a larger total surface area, which will in turn contribute to more efficient electron transfer. According to the XRD data (see
Table S2), the crystal size of the NPs decreased in the order of
T2 >
T3 >
T1, and the recorded current also increased in the manner of
T2 <
T3 <
T1. On the other hand, the ΔE of the first redox couple indicated good reversibility for
T1 and
T3 compared to
T2. However, the second redox couple was more reversible in
T2, which also demonstrated the symmetry of the peak shapes.
In all cases, but especially in the case of
T1 and
T3, the peaks of the first redox wave were higher than those in the second wave. In
Cb,
C1,
V1 and
T1,
T2,
T3, the alkali atoms Na and K, respectively, were intercalated into the vacancies of the PB crystal lattice. Additional K
+ from the electrolyte can intercalate into the lattice when reducing PB to PW, while the alkali atoms can be removed from the crystal lattice when oxidising PB to BG. Accordingly, in all particles, the first redox couple may be more efficiently accessible to the electrolyte, leading to faster diffusion and higher current, as shown in
Figure 6a,b.
The POD-mimicking catalytic activity of the PBNPs and PB/PEI NPs was investigated by incubating them with a chromogenic ready-to-use (TMB)/H
2O
2 substrate solution. All nanoparticles exhibited POD-mimicking catalytic activity, reducing H
2O
2 to H
2O and oxidising the colourless TMB
Red to its blue redox intermediate TMB
Red/TMB
Ox, as HRP does (
Figure 7a) [
45]. The reaction was then stopped by adding sulphuric acid, which produced yellow fully oxidised TMB
Ox, which was measured at 450 nm.
Figure 7b presents the signals measured as a function of nanoparticle type and concentration up to 12.5 µg mL
−1 (signal saturation was reached at higher concentrations for all nanoparticles).
Figure 7c depicts the limit of detection (LOD) for each curve, which indicates the lowest concentration of nanoparticle that could be detected. At first glance, it was evident that all PB/PEI NPs except
T2 produced higher signals than
Cb, with higher catalytic activity correlating with lower LOD. However,
Cb displayed a lower LOD than most PB/PEI NPs (27 ng/mL). This was attributed to PEI, which could alter or shield the PB catalytic sites, making it harder to detect low particle concentrations but providing protonated amine groups that could also facilitate electron transfer, especially at high particle concentrations.
When the two types of nanoparticles produced by coprecipitation were compared,
C1 and
Cb had an approximately similar size, however,
C1 exhibited a higher catalytic activity but slightly higher LOD.
C1 showed a lower reversibility of the first redox wave in the CV and lower colloidal stability, thus a lower exposure of active surface to the medium (zeta potential 0.07 ± 2.5 mV and −46.8 ± 6.5 mV for
C1 and
Cb;
Figure S2). The EDX results indicated that
C1 displayed slightly higher Fe to alkali content (88.63% and 85.06% Fe for
C1 and
Cb, respectively; see
Table S1), which could result in more catalytic sites, while the Fe
2+/Fe
3+ ratio was lower (0.88 versus 0.94). The main difference seemed to be the PEI presence in
C1, with an estimated 5.3% of nitrogen derived from the PEI, as determined by XPS.
On the other hand,
V1 displayed the highest catalytic activity and lowest LOD observed. It was anticipated that smaller particle sizes should result in enhanced catalytic activity due to the increased surface area [
46].
V1 was the smallest nanoparticle tested (19 ± 4 nm apparent size and 23 nm crystalline size;
Table S2) and was also characterised by high colloidal stability (zeta potential 38.0 ± 4.4 mV) and the highest microstrain (0.9%), Fe to alkali content (94% Fe), Fe
2+/Fe
3+ ratio (1.37), and PEI incorporation (25.3%). Thus, the reason for the high catalytic activity of
V1 was presumably the combination of small particle size, the increased amount of well exposed active surface, elevated number of highly active catalytic sites, and efficient PEI-driven wiring.
For the particles produced via thermal reduction, T3 demonstrated higher catalytic activity than T1, and T2 exhibited the lowest catalytic activity among all particles. Again, smaller particles worked better. The bad performance of T2 was correlated to the low peak current and reversibility observed by CV. Of these three types of particles, T2 also displayed the lowest Fe content, but the numbers were still slightly larger than those of Cb, which performed better. Furthermore, the concentration of PEI in T2 was between those in T1 and T3.
Globally, the results suggest that none of the parameters studied was solely responsible for nanoparticle catalytic activity. However, a combination of small size, high iron content, and high Fe2+/Fe3+ ratio seemed determinant for particles displaying high catalytic activity. The incorporation of the aminated polymer could facilitate electron transfer between the nanocatalyst and the TMB/H2O2 system in some cases, but also disrupt or shield the PB catalytic active sites in others. Finally, while, in general, bad catalysts produced distorted CVs and good catalysts produced larger and more reversible peaks, peak height and reversibility were not useful to solely define catalyst efficiency. Independently of this, the LODs calculated for all of the nanoparticles were higher than those obtained for HRP, which was a better catalyst.
The steady-state kinetic behaviour of the nanoparticles was investigated further, this time using a fixed concentration of TMB and increasing concentrations of H
2O
2 (
Figure 8). The Michaelis constant (Km) and maximal velocity (Vmax) were calculated by utilising the Michaelis–Menten model (
Figure 8a–f) and the Lineweaver–Burk reciprocal model (
Figure 8a’–f’), respectively. The Km values were found to be 2.66, 3.35, 4.10, 4.39, and 6.03 mM for
T1,
T3,
C1,
V1, and
Cb, respectively, while
T2 exhibited a higher value of 85.19 mM (see
Table S5). Since Km defines the affinity between H
2O
2 and the catalyst [
47], the values obtained revealed a higher H
2O
2 affinity towards PB/PEI NPs than to
Cb, with the exception of
T2. Although the highest affinity was obtained in the low particle size range (<50 nm), the smallest
V1 displayed an unexpected low affinity for H
2O
2. On the other hand, the Vmax was approximately of the same order for all types of nanoparticles (4.76 × 10
−5 mM s
−1, 4.35 × 10
−5 mM s
−1, 4.30 × 10
−5 mM s
−1, 4.29 × 10
−5 mM s
−1, 4.11 × 10
−5 mM s
−1, 3.38 × 10
−5 mM s
−1 for
T3,
T2,
C1,
Cb,
V1, and
T1, respectively). While Km and Vmax are frequently employed to evaluate the catalytic activity of PBNPs, the LOD value provides a simpler and more direct measure of their catalytic performance (
Table S5). The
V1,
T1, and
T3 nanoparticles demonstrated the optimal POD-like catalytic performance across the evaluated parameters, along with the lowest LOD, the lowest Km, and the highest Vmax, respectively.
The catalytic activity of the PBNPs can be predominantly attributed to the redox cycling of the oxidative and reductive states of PB, coupled with electron transfer between the PB and TMB/H
2O
2. As proposed by Feng et al., in our system, PB could undergo a valence band pathway by transferring an electron to reduce H
2O
2 (causing PB oxidation), then accepting an electron from TMB (oxidising TMB and reducing oxidised PB). An alternative hypothesis is that the process of the conduction band pathway occurs when an electron is transferred from TMB to PB or pre-oxidised PB (causing the reduction of PB or pre-oxidised PB). This is followed by the release of an electron, which reduces H
2O
2, while oxidising PB or reduced state PB. Furthermore, as hypothesised by Komkova et al., PB initially reacts with TMB by reducing to PW and oxidising TMB, then reduces the H
2O
2 by oxidising PW [
48,
49].
Conversely, the iron valence states of PB have the potential to generate reactive oxygen species (ROS), which can influence the catalytic reaction, particularly in particle modifications and/or with external energy. However, it should be noted that PB is also capable of scavenging ROS [
15,
48,
50]. Nevertheless, understanding the precise mechanism is challenging because of the complexity and incomplete elucidation of its various operational processes. Moreover, it is noteworthy to mention that the chromogenic agent TMB suffers from a lack of specificity to be applied in deep studies of the POD mechanisms [
51].
It is well-known that HRP activity is impaired in the presence of a number of inhibitors [
52,
53,
54]. The POD-like activity of all synthesised particles was evaluated in the presence of a range of HRP inhibitors including NaN
3, H
2SO
4, H
2O
2, and EDTA.
Figure 9 depicts the evolution of the blue colour resulting from the oxidation of TMB, which was catalysed by the HRP or nanozymatic PBNPs (
Cb and
C1) at varying concentrations in the presence of the aforementioned inhibitors at different concentrations.
Both
Cb and
C1 exhibited highly stable catalytic behaviour in the presence of 1.92–6.79 M NaN
3 in comparison to the extreme inhibition caused to HRP (
Figure 9a). The same was observed for
T1,
T2,
T3, and
V1 (
Figure S3), without any impact of the POD-like activity. This indicates that the coordination complex of PB, or its redox Fe
2+/Fe
3+system, is not influenced by NaN
3. The catalytic activity and kinetics of HRP are also highly sensitive to acidic pH, which alters the enzyme structure and deactivates its active sites (
Figure 9b-HRP) [
18]. However, POD-like deactivation was not observed for the PBNPs and PB/PEI NPs (
Figure 9b and
Figure S4). On the contrary, when stored in basic solutions these particles tend to degrade, and the rate of degradation depends on the strength of the basicity and the storage time.
Figure 9c illustrates the gradual reduction in the catalytic activity of HRP observed when the TMB/H
2O
2 substrate solution was supplemented with increasing H
2O
2 concentrations of up to 1 M. This can be attributed to the formation of inactive enzymatic species, E
x, leading to a reduction in catalytically active enzymatic sites [
52]. It is noteworthy that for
Cb and
C1, as illustrated in
Figure 9c, the catalytic activity was not only retained, but increased proportionally to the H
2O
2 concentration (see also
Figure S5). This implies that the commercial TMB/H
2O
2 system, which is produced for HRP detection, includes a concentration of H
2O
2 far from optimal for PB maximal efficiency.
In contrast, the
Cb bare PBNPs demonstrated higher catalytic inhibition by EDTA than HRP (
Figure 9d). This was particularly evident at particle concentrations below 25 ng/mL in high EDTA levels. EDTA is a potent metal chelating that binds to Fe
2+ and Fe
3+ centres through its amine and carboxylic groups, forming complexes [
55]. This results in the reduction of redox-active catalytic metal sites and POD-like activity in the PBNPs. Nevertheless, the PEI-modified particles
C1,
T1, and
V1 (see
Figure S6) exhibited a lower impact on their activities. This could be due to the presence of the aminated PEI, which may repel EDTA to a certain extent, reducing the chelation of the metal centres.
Globally, these results show that PBNPs and PB/PEI NPs are more resistant than HRP to at least a number of potential inhibitors.