Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 14 September 2023
Sec. Photocatalysis and Photochemistry
This article is part of the Research Topic Recent Advances in Photokinetics View all 4 articles

On photokinetics under monochromatic light

  • Leicester School of Pharmacy, De Montfort University, Leicester, United Kingdom

The properties of photokinetics under monochromatic light have not yet been fully described in the literature. In addition, for the last 120 years or so, explicit, handy model equations that can map out the kinetic behaviour of photoreactions have been lacking. These gaps in the knowledge are addressed in the present paper. Several general features of such photokinetics were investigated, including the effects of initial reactant concentration, the presence of spectator molecules, and radiation intensity. A unique equation, standing for a pseudo-integrated rate law, capable of outlining the kinetic behaviour of any photoreaction is proposed. In addition, a method that solves for quantum yields and absorption coefficients of all species of a given photoreaction is detailed. A metric (the initial velocity) has been adopted, and its reliability for the quantification of several effects was proven by theoretical derivation, Runge–Kutta numerical integration calculations and through the model equation proposed. Overall, this study shows that, under monochromatic light, photoreaction kinetics is well described by Φ-order kinetics, which is embodied by a unifying model equation. This paper is aimed at contributing to rationalising photokinetics via reliable, easy-to-use mathematical tools.

1 Introduction

Light-sensitive chemicals are ubiquitous in nature and increasingly becoming a part of new or potential technological and industrial processes. From a general application point of view, two main themes stir the interest in such molecular systems and materials: relating to their direct potential benefit and/or to avoid their degradation. For instance, photolabile molecular systems find applications in nanotechnology (Shit et al., 2020), biomimetic chemistry (Cahova et al., 2013), optical information storage (Irie, 2015; Nakatani et al., 2016; Sarter, et al., 2016; Pianowski, 2022), surface science (Bonacchi et al., 2015), environmental chemistry (Nakatani et al., 2016; Malato et al., 2021), photo-responsive materials (Irie, 2000; Irie et al., 2014; Okuda et al., 2016), photoswitchable nucleosides (Singer and Jaschke, 2010; Cahova and Jaschke, 2013; Singer et al., 2013; Chen et al., 2014), pharmaceutical drugs (Velema et al., 2014), photoenergy (Montalti et al., 2020), and green chemistry (Scaiano, 2020; Pianowski, 2022), to cite a few fields. The trend of such applications is expected to grow in the future (Bonfield et al., 2020).

One of the primary tools for the study of molecular reactions is chemical kinetics, specifically photokinetics in the case of photoreactions. Among its aims, photokinetics allows to unravel the intrinsic parameters (i.e., quantum yields and absorptivities) of the reaction at hand. Such information is necessary for the understanding of reaction behaviour and control. Unfortunately, the area of photokinetics remains significantly underdeveloped compared to (thermal) chemical kinetics. The former not only lacks the usual metrics used in chemical kinetics (such as the kinetic order of a reaction and/or integrated solutions of the rate laws) but also suffers from the absence of standard methods, a situation that might explain a considerable lack of text books on the subject, where, surprisingly, we only find a single entry that is fully dedicated to photokinetics (a textbook published in 1998) (Mauser and Gauglitz, 1998).

The main hindrance facing a normal development of photokinetics can be traced down to a mathematical impossibility to analytically solve the rate laws of photoreactions (Mauser et al., 1998; Crano and Guglielmetti, 2003; Tonnesen, 2004; Nakatani et al., 2016; Montalti et al., 2020; Scaiano, 2020; Pianowski, 2022). The absence of integrated rate laws has considerably limited photokinetic investigation towards unravelling photoreaction features, offering accessible metric tools to the experimentalist, and/or allowing better control of the reactions, as usually available for thermal kinetic studies.

In this context, several approaches have been usually adopted to circumvent this mathematical hurdle. In the first instance, one can notice the ubiquity of the usage of thermal kinetics’ equations (such as those corresponding to zeroth- and first-order kinetics) in the literature that deals with data of photoreactions (Crano and Guglielmetti, 2003; Tonnesen, 2004; Nakatani et al., 2016; Malato et al., 2021; Pianowski, 2022). Despite its popularity, this approach can be deemed inappropriate due to a significant difference between the mathematical formulations of the rate laws describing thermal and photochemical reactions.

A second, more elaborate method employs expansion with a power series of the rate-law terms involving power numbers, in order to achieve integration of the rate law. In general, this approach is applied to the simplest photoreactions (e.g., no known example has treated the photoreversible reaction) and limits the expansion to the first order of the power series. Even if an analytical solution can be achieved in this case, it still remains of limited application due to the stringent condition imposed on the allowed magnitude of the absorbance (which might be as low as 0.01). If expansion to a higher order of the power series is envisaged, in order to alleviate the absorbance limit and reach workable experimental values, the differential equation becomes, unfortunately, unsolvable analytically.

Till now, the best treatment of photokinetic data has been achieved by employing numerical integration methods (NIMs). NIMs are very powerful tools for this purpose, but their usage and technicalities are not necessarily familiar to most experimentalists in the field of photokinetics. More importantly, the efficacity of NIMs remains dependent on the number of unknown reaction parameters to define. In general, NIMs work fine when only a small number of parameters are sought but will face an identifiability problem (Maafi and Brown, 2005a) as the number of those parameters increases (for instance, NIMs are not capable of unravelling the true parameters for a photoreversible reaction if both forward and reverse quantum yields, as well as the absorption coefficient of the photoproduct, are unknown). We shall discuss this issue further in Section 3.5.

According to the aforementioned succinct review, photokinetics is still in need of both i) a simple way that should be accessible to all, allowing photokinetic data investigation, and ii) a standard and reliable set of model kinetic equations that allow consistent photokinetic studies and conclusions.

In our team, we have previously proven that the photokinetics of the primary photoprocess (XY, whose photoproduct, Y, is transparent to the monochromatic irradiating beam) can be solved analytically and the reaction’s behaviour obeys Φ-order kinetics (Maafi and Brown, 2007). The analytically derived solution for this reaction provides both an integrated rate law involving a logarithmic function bearing a time-exponential term in its argument (vide infra Eq. 4) and an explicit formula for the rate constant of the photoreaction (Maafi and Brown, 2007). However, when the product of the primary photoprocess absorbs concomitantly with the reactant, there are two distinct kinetic cases: a) the rate law of the reaction becomes non-integrable if the monochromatic light wavelength is different from that of an isosbestic point (non-isosbestic irradiation) (Maafi and Maafi, 2013) and b) the reaction obeys the first-order kinetics when the irradiation light is isosbestic (Maafi and Brown, 2005b). It is interesting to notice that such a dual (a/b) kinetic pattern was preserved for all sub-mechanisms of a cyclic trimolecular system involving six photochemical steps (Maafi and Brown, 2005b).

Furthermore, a semi-empirical approach combining Runge–Kutta (RK) numerical integration trace data and an explicit Ct algebraic equation of the Φ-order type was proposed to overcome the latter mathematical hurdle, when irradiation is non-isosbestic, and deliver exploitable explicit models for other reactions than the primary photoreaction discussed previously. The aim was to achieve consistency and precision in describing the photokinetics and to determine both unknown quantum yields and absorption coefficients of the reaction at hand. The approach was developed and experimentally applied to the primary photoprocess with absorbing the reactant and product (Maafi and Maafi, 2013), the photoreversible reaction (Maafi and Maafi, 2014a; 2014b), and the multi-consecutive photoreaction involving four photoproducts (Maafi and Maafi, 2016). One of the interesting experimental results facilitated by those formulae was evidence of the variability of the quantum yield (of each individual reaction step) with the wavelength of the irradiation light, as has been shown for a variety of photoreactive molecules that belonged to different chemical families (Maafi and Brown, 2007; Maafi and Maafi, 2013; Maafi and Maafi, 2014b; Maafi and Lee, 2015a; Maafi and Lee, 2015b; Maafi and Maafi, 2016; Maafi and Al qarni, 2019; Maafi and Al qarni, 2022).

Even if the Ct explicit expressions proposed for the latter reactions were very useful, and the semi-empirical method can, in principle, be developed/applied to any photoreaction, it is, however, a fact that the semi-empirical equations are, thus far, only available for a handful of reactions, and they are restricted to absorbance limits.

Nonetheless, our results have proven that using explicit formulae is not only handy but also provides a better assessment of reaction photokinetics. However, there is a gap in the knowledge that needs to be addressed. From this perspective, it would be an advantage to have a model equation that fits all reactions without imposing any constraints on the initial concentrations or absorbance. It would also be of great interest to have a methodology and metrics to assess the reaction, define its unknown quantum yields and absorption coefficients, evaluate the quantum yield variation with wavelength, and develop new actinometers. The present study is dedicated to an attempt to bring adequate answers to the aforementioned points for photoreactions performed under monochromatic light. In addition, it proposes a review of the main properties of the reaction considered in different situations. Such an overview has not been presented, thus far, in photochemistry and kinetic literature.

2 Experimental

Each tested reaction mechanism is identified by a reactant X, j photoproducts (Yj) (where 0jnsp, with X=Y0, and nsp is the total number of species in the reaction medium), and nΦj photoreaction steps starting or ending at species Yj, according to the Φ-shaped mechanism given in Scheme 1.

Time evolution of concentrations corresponding to each species involved in the mechanism (kinetic traces CXt and CYjt) was numerically calculated. The NIM selected for the present work consisted of a fourth-order RK NIM. The RK-calculated data were generated by a homemade programme. The code runs on an Excel VBA platform that is available on Microsoft Excel.

RK-calculated traces served as a reference for testing the performance of the proposed model formula (vide infra Eq. 4). The RK simulations were conducted at iteration intervals ranging between 0.1 and 5 s. The fitting of the RK-generated traces to Eq. 4 (for X and each Yj of a given mechanism) was performed with a Levenberg–Marquardt algorithm (LMA) provided by the curve-fitting tool of R2020b Matlab software. The goodness of fit of the traces was assessed by i) the values of the squared correlation coefficients (r2), characterising the linear plot between the data of the RK-simulated traces and those supplied by the LMA-calculated curves based on the respective Eq. 4 of the reactant and each species Yj of the reactive system, ii) the sum of squares error (SSE), and iii) the root mean square deviation (RMSD) between the two datasets. The parameters of Eq. 4 (wij, ccj, and kij) were obtained by LMA within 95% confidence limits.

The number of mono-Φ-order terms (i.e., ωLog1+ccekt) in each Eq. 4, describing each of the traces of the species involved in the reaction mechanism at hand, cannot exceed the maximum value of nΦ, but it can be lower than nΦ (i.e., ij), depending on the trace. The RK calculations assume de facto that the concentrations of the reactant and the Yj species belong to the respective linearity ranges of those species’ calibration graphs, throughout the duration of the reaction (irrespective of the reaction investigated) (it is to be noted that Eq. (1) is not valid for high concentrations that lay beyond the respective linearity ranges of the species).

The nsp fitting equations (of Eq. 4 type) corresponding to the reactant and j photoproducts (involved in a given reactive system) are effectively coupled in the sense that Eq. 4 of species Yj necessarily shares some, if not all, of its k values with Eq. 4 of previous species Yj1 (occurring in the reaction mechanism before species Yj).

The rate laws considered in the present work apply to a slab-shaped, continuously and vigorously stirred reactor subjected to a collimated monochromatic light beam.

3 Results and discussion

3.1 The rate-law equation

As for thermal kinetics, in photokinetics, a rate law should be written for each species Yj (where 0j7, with the reactant X= Y0 and the photoproducts Y1to7) of the reaction mechanism (Scheme 1). The nΦj reaction steps linking Yj can correspond to either forward or backward reactions with other Yj species (jj). Each rate law, for a given Yj (rYjλirrt), can then take the following general formulation, irrespective of the actual reaction mechanism.

rYjλirrt=CYjλirrt=jjnΦjΦYjYjλirrPaYjλirrt+ΦYjYjλirrPaYjλirrt.(1)

SCHEME 1
www.frontiersin.org

SCHEME 1. Φ-shaped photomechanism involving eight reactive species interlinked by 14 photochemical reaction steps.

It is to be noted that it is unusual in kinetics that the concentration of species Yj at time t is written relative to the irradiation wavelength, as given in Eq. (1). This labelling is necessary simply because all the parameters on the right-hand side of Eq. (1) are dependent on λirr. This makes two traces of a species, obtained at two different wavelengths λirr1 and λirr2 (e.g., CYjλirr1t and CYjλirr2), most certainly different from each other.

In Eq. (1), the light absorbed by species Yj or Yj (PaYjλirr and PaYjλirr, with jj), constitutes fractions of the total absorbed light by the medium (Paλirr) at time t (expressed in einsteins1dm3), according to the following expression:

PaYjorjλirrt=AYjorjλirrtAtotλirrtP0λirr110Atotλirrt=AYjorjλirrtP0λirrPKFt,(2)

where P0λirr is the intensity of the incident radiation (see Section 3.9 for a detailed definition), the total absorbance (Atotλirr) is a sum of the individual absorbances of the species (AYjλirr) at λirr (nsp, in Eq. (3), represents the number of all species Yj—reactant and products—in the medium at time t), and PKFt is the dimensionless photokinetic factor.

Atotλirrt=j=0nspAYjλirrt.(3)

Obtaining analytical solutions of Eq. (1) rate-law equations, for a vast majority of cases, is simply impossible. This mathematical hurdle can be explained by the fact that these differential equations are non-linear due to the presence of the time-variable term 110Atotλirr, which encompasses all absorbances of the species present in the reactive medium. This situation is expected to persist for the foreseeable future until a mathematical technique, capable of analytically solving these complicated differential equations, is devised.

3.2 The proposed integrated rate-law model

The traces of a primary photoprocess subjected to monochromatic light were proven to obey a Φ-order equation that was analytically derived from the reaction’s rate law (Maafi and Brown, 2007). Previous work on other photomechanisms (Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2016), whose rate laws cannot be solved by closed-form integrations, has also shown that the traces of photoreactions, involving up to five photochemical steps, can be well described by specific semi-empirical formulae, which mathematically possess a Φ-order character, but under certain constraints on the total absorbance at the end of the reaction (Atotλirr ranging between 0.2 and 0.6). However, till now, no general formula has been proposed for the description of traces of photoreactions. A conjecture, based on the aforementioned results, might suggest that such a general equation would be of a Φ-order type.

Accordingly, the following explicit formula (Eq. 4) is proposed to describe the trace of species Yj (CYjλirrt, j=0 for the reactant X), irrespective of the mechanism undergone by the actual photoreaction, and the wavelength (λirr) of the non-isosbestic monochromatic irradiation light driving the reaction. The reaction mechanism may involve any number of photochemical reaction steps (e.g., within Scheme 1). The temporal variation in the concentration of a typical species Yj is given by Eq. 4 (where Log and e are, respectively, the base 10 logarithm and the exponential functions and ij is the number of mono-Φ-order terms under the sum).

CYjλirrt=C,jλirr+i=1ijωijλirrLog1+ccjλirrekijλirrt.(4)

The expression for the reaction rate of species Yj is obtained by the differentiation of the corresponding Eq. 4 as

CYjλirrt=rYjλirrt=i=1ijωijλirrccjλirrkijλirrekijλirrt1+ccjλirrekijλirrtln10,(5)

from which we can derive the formula for Yj initial reaction rates (expressed in Ms1) as

CYjλirrtt=0=r0Yjλirr=1ln10i=1ijωijλirrkijλirrccjλirr1+ccjλirr.(6)

The parameters (C,jλirr, ωijλirr, ccjλirr, and kijλirr) of Eq. 4 are specific to the irradiation wavelength (λirr), making CYjλirr wavelength specific. The parameter C,jλirr, standing for the final concentration of species Yj, will be non-zero, but positive, if the species persists at the end of the reaction (typically for photostationary reactions and end products). The pre-exponential coefficient ccjλirr is considered here to be the same for all ij terms making up Eq. 4 of a given species Yj (ccjλirr=ccλirr, a constraint that is verified practically). The pre-logarithmic parameter ωijλirr, accepting positive and negative values, might be considered a weighing factor for the logarithm of the ith kinetic regime of the considered species Yj. Parameter kijλirr is the positive rate constant of the ith kinetic regime (expressed in s1) and is proportional to the light intensity and the quantum yield of the ith kinetic regime, but it may also depend on other attributes of the reactive species (these details are currently unknown).

3.3 The mechanism selected

In order to test the proposed equations (e.g., Eq. 4), a Φ-shaped mechanism is selected (Scheme 1, involving the reactant, X, seven photoproducts, Y, and 14 reaction steps). It is thought to be representative of a number of simple and complicated photoreactions encountered in photochemistry books and literature (Mauser et al., 1998; Crano and Guglielmetti, 2003; Tonnesen, 2004; Nakatani et al., 2016; Montalti et al., 2020; Scaiano, 2020; Pianowski, 2022). The number of sub-mechanisms that can be worked out from Scheme 1 exceeds 100 examplars (including the counting of the absorption of terminal species that might equal to zero).

The set of sub-mechanisms worked out from Scheme 1 includes, but is not limited to, those of the primary photoprocess with a transparent photoproduct (as for photochromic diarylethenes under visible light (Crano and Guglielmetti, 2003; Nakatani et al., 2016; Maafi and Alqarni, 2022; Pianowski, 2022)), the primary photoprocess with an absorbing photoproduct (as for nifedipine, nisoldipine, and dacarbazine drugs (Tonnesen, 2004; Maafi and Maafi, 2013; Maafi and Lee, 2015a)), the photoreversible reaction (as for stilbenoids (Crano and Guglielmetti, 2003; Maafi and Al-Qarni, 2019), anti-cancer drugs (Tonnesen, 2004; Maafi and Lee, 2015b), and dimethylhydropropenes (Ziani et al., 2021)), the consecutive double photoreversible system (as for fulgides in solution or in the solid state (Crano and Guglielmetti, 2003; Weerasekara et al., 2017; Kochman et al., 2022; Pianowski, 2022) and benzopyrans (Frigoli et al., 2020)), the multi-consecutive photoreaction (as for riboflavin (Maafi and Maafi, 2016)), and other more complex systems (Nakatani et al., 2016; Scaiano, 2020; Malato et al., 2021; Pianowski, 2022). The mechanisms and molecular systems reviewed previously have been extensively investigated, but their photokinetic studies tended to be either performed by using the equations of thermal kinetics (e.g., first-order exponential models), using various proposed (non-analytical) equations, or employing numerical integration. It is to be noted that no standard equation model has ever been provided in the aforementioned documentation and no integrated rate law has ever been analytically derived for photoreactions (including the simplest ones, such as the concurrent Y3XY1 reaction).

3.4 Some general aspects of photo- and thermal kinetics

From a kinetic view point, the rate of a thermal reaction of a species Yj does not depend on the properties (e.g., concentrations) of distant species from Yj (that are not generated by or generate themselves Yj), whereas in photokinetics, reactions’ rates depend not only on the properties of the species involved in the considered reaction step, as in thermal kinetics, but also on those of all the absorbing species occurring in the reaction medium over time. In fact, the rate of a photochemical reaction depends explicitly on the absorbances of all the species with a non-zero absorbance at a given time t. This particular difference between photo- and thermal kinetics is expressed in the rate equation (Eqs 1, 2) by the total absorbance of the set of reactive species (Atotλirrt) present in the medium at a given time. Such a subtle interdependence of the reaction species within the photokinetic rate law is responsible for the species’ competition for the incident light. Therefore, species that belong to completely separated reaction steps of the overall reaction mechanism still influence each other’s reactivities. In the same context, the kinetics of species Yj might also be dependent on the absorbance of one or more inert (non-reactive) species, present in the medium as spectators, when the latter absorb at the irradiation wavelength. This particular phenomenon is further described in Section 3.7.

Furthermore, since the coefficients of the rate-law equation correspond to the attributes of the species and light at a given wavelength, it is evident that the solution of that equation is also specific to that wavelength. Therefore, changing the irradiation wavelength will cause the trace to change accordingly. This imposes that all Eq. 4 coefficients are indexed by λirr (including the concentrations, CYjλirrt). In thermal kinetics, this would be equivalent to a change in temperature. Here, apart from a qualitative comparison (e.g., reaction speeds up between irradiations at λirr1 and λirr2), it is difficult to proceed with further quantitative analysis based on the comparison of the traces in the absence of an integrated rate law and the quantitative parameters it provides. For instance, the knowledge of the relative reaction velocity at a given time but at different wavelengths can help selecting the light wavelength range causing the species at hand to be most or least reactive depending on the purpose of the experimentalist/experiment. For some molecules (drugs, photochromes, etc.), it is it useful to specify the regions of light inducing most photoconversion, a situation which would be solved for photoreactions using Eq. 4, as will be shown in the following sections. One additional difference between thermal and photo-kinetics concerns an evidenced change in the photoreaction kinetic order. It was previously proven that the trimolecular cyclic photoreaction (Maafi and Brown, 2005b) and all its sub-mechanisms obeyed a first-order kinetics when irradiated with an isosbestic monochromatic beam (λisos coincides with an isosbestic point on overlayed successive absorption spectra of the medium through reaction time). It is reasonable to predict that any photoreaction will also be described by the first order kinetics whenever subjected to an isosbestic irradiation since the photokinetic factor [PKFt, Eq. (2)], under these conditions, is independent of time (which converts the rate law into a first-order linear differential equation). However, the kinetics of any photoreaction, under non-isosbestic light, extracted from the Φ-shaped mechanism, is shown in the following sections to follow Φ-order kinetics. The conjecture is to observe similar conclusions for any photoreaction. Accordingly, a general statement can be derived from the aforementioned observations: the kinetic data of a photoreaction can either obey a first-order or a Φ-order kinetics, respectively, when driven by an isosbestic or a non-isosbestic irradiation light (when all other experimental and reaction attributes are kept the same). A true change in the kinetic order then specifically occurs for photoreactions (the equivalent for thermal reactions would perhaps correspond to a change in the kinetic order when temperature changes, which is not observed). Also, it is noticeable that the mathematical formulation of the rate law explicitly includes a rate-constant factor (k) for thermal reactions but not for photochemical reactions.

Finally, the concentrations of the species involved in a photokinetic process must all belong to their respective calibration graphs (a redundant condition for thermal kinetics).

3.5 Total absorbance and fitting RK traces with the model equation

For generality purposes of the application of the model equation (Eq. 4), RK simulations were conducted on several sub-mechanisms derived from Scheme 1. Individual RK traces have been obtained for each species, involved in the sub-mechanism considered, and the fitting of each of these traces was performed with its corresponding Eq. 4.

The process of RK trace fitting was implemented according to a simple protocol. The first trace to be fitted was the reactant’s, followed by that of the nearest Yj species in the reaction mechanism, then the next, and so on. The fitting parameters kijλirr obtained for an earlier species (e.g., k10λirr and k20λirr for X and j=0) were kept the same in the subsequent equations, describing the trace of the next species linked to X, where the latter might involve new terms (e.g., k11λirr=k10λirr, k21λirr=k20λirr, and k31λirr for Y1). Kinetically, this means that the trace of a given species in the reaction mechanism is necessarily dependent on the rate constants (both their number and values) of previous reaction steps generating that species in the reaction mechanism. Accordingly, the number ij of mono-Φ-order terms in Eq. 4 is minimal for the reactant and is expected to increase (or at least remains the same) for species occurring subsequently. The maximum number of terms in Eq. 4 that describe the trace of any single species cannot exceed the number of reaction steps occurring in the considered overall reaction mechanism (ijnΦ). The minimum number of these terms should, in principle, be equal to the number of reaction steps starting or ending at the considered species (i.e., nΦj). However, in some cases and depending on the kinetics, Eq. 4 can provide a good fit of the traces with less than nΦj terms. Interestingly, it is predicted that most likely the same number of terms is found for each Eq. 4 of each species in a cyclic reaction mechanism involving three or more molecules (e.g., Figure 1).

FIGURE 1
www.frontiersin.org

FIGURE 1. Example of a good fit of the reaction’s species RK traces (circles) by their corresponding Eqs. 4 (lines) for the indicated cyclic reaction mechanism. Each Eq. 4 of each of the species counted four mono-Φ-order terms, a unique set of four kijλirr values, and a single common value of ccλirr.

Furthermore, it has been observed that this fitting protocol indicates that factor ccjλirr keeps the same value (ccjλirr=ccλirr) for all Eq. 4 of all the species belonging to a given reaction mechanism. C,jλirr takes the value of the final concentration of the considered species Yj, and ωijλirr are different for different traces of the species of the same mechanism.

This coefficient, ccλirr, hence, plays the role of a coupling factor of the system of Eq. 4, describing the studied reaction. This feature will impose an additional constraint on the fitting of the photokinetic traces, which is expected to improve the reliability of the fitting parameters’ values.

It is possible to add a few more constraints to the fitting protocol, by considering Eq. 4 at the initial time (t=0) and the known initial concentration values of the species.

CYjλirr0=C,jλirr+i=1ijωijλirrLog1+ccjλirr.(7)

In the present study, more than 200 RK traces were fitted according to the protocol described previously. These traces belonged to species involved in various sub-mechanisms derived from the Φ-shaped reaction (Scheme 1), including one to seven photoproducts and 1 to 14 reaction steps.

The individual species’ RK-generated traces were fit by adequate Eq. 4 of the reaction studied with various numbers of mono-Φ-order terms making up each of these equations. The protocol described previously for the treatment of the RK data with Eq. 4 worked extremely well for all traces. Excellent fittings of each individual trace (illustration in Figure 1) were obtained with correlation coefficient values, for RK-calculated vs. Eq. 4 data, not lower than 0.999, sums of squared errors (SSE) as low as 10−22, and root mean square errors (RMSE) of no higher than 10−9.

These results prove that the Φ-order kinetic behaviour is preserved for photochemical reactions, irrespective of the operating sub-mechanism that can be extracted from Scheme 1. This stands for a confirmation of our conjecture that the Φ-order equation is the seminal model for photoreactions despite the fact that it has only been analytically derived for the primary photoprocess whose photoproduct is transparent to the radiation used (Maafi and Brown, 2007). Such a situation might be an analogue to that observed for thermal reactions where the mono-exponential function, obeyed by the kinetics of the simplest unimolecular reaction, is the basis for the description of kinetic data belonging to any other, more complex thermal reaction mechanisms. Incidentally, this is somewhat supported by the typical Φ-order equations set out for Ct by the semi-empirical method for several reactions (Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2014b; Maafi and Maafi, 2016).

The findings of the present work also allow us to put forward a conjecture stipulating that Φ-order kinetics (based on Eq. 4 template) maps out, in general, the behaviour of photoreaction kinetics (irrespective of the photomechanism considered, i.e., beyond Scheme 1).

Hence, the formulation of the integrated rate equation given by the proposed Eq. 4 becomes the unifying model in photokinetics. This represents the first time that a unique explicit description is proposed for the traces of photoreactions.

Conversely and even if some photokinetic traces may well be fitted by one or several mono-exponential functions, such an approach remains limited and cannot be generalised. It is unlikely to reach similar fitting results by using mono-exponential functions instead of Φ-order equations (of Eq. 4 template), keeping the kinetic meaning of the fitting results. It has already been shown that such an interchange of (Φ-order/exponential) equations does not always work for the simplest primary photoreaction (Maafi and Brown, 2007). More importantly, this observation is fully justified, in principle, by the significant difference between the mathematical formulations of the rate laws describing photoreactions and thermal reactions. These differences in the differential equations lead to a predictable dissimilarity of the mathematical solutions that would be derived for each (e.g., first- and Φ-order kinetics).

In addition to the successful fitting of the traces of photoreactions by the unifying Eq. 4 model, this study has revealed an aspect that must be considered in any practical photokinetic investigation. Indeed, despite the constraints imposed on the fitting of the traces [as given by the coupling factor ccλirr and Eq. (7)], it has become evident that reaching a unique solution (a unique set of values for the fitting parameters ccλirr, ωijλirr, and kijλirr) for a given reaction is not achievable. This translates the occurrence of an identifiability issue (Vajda and Rabitz, 1988; Maafi and Brown, 2005; Hattersley et al., 2011; Ovchinnikov et al., 2023), i.e., changing the initial values where the fitting calculation starts, yields a quite large number of different sets of fitting values of the parameters. Each of these sets has equally excellent fitting metrics of the RK traces, indicating that the Levenberg–Marquardt algorithm can converge to different minima. The reason behind the identifiability issue is a smaller number of linearly independent equations describing the reactive system compared to the number of fitting parameters. In addition, Eq. (7) is linearly independent from Eq. 4, but is non-linear.

This point is well illustrated in the case of the simplest photoreaction (XY1, whose photoproduct is transparent to the irradiation light). Here, the trace is described by a mono-Φ-order equation. Nonetheless, it is not possible to single out the true solution of the system (defined by the true values of ω10λirr, ccλirr, and k10λirr describing the reaction), where successive fittings (with different initial values) yield several possible sets of fitting parameters’ values with excellent fit of the trace. For more extended reactions, the insolvability of the identifiability issue becomes obviously more acute as the ratio between parameters’ and equations’ numbers becomes even larger. Overall, the fundamental problem posed here emerges from the unavailability of ωijλirr, ccλirr, and kijλirr explicit expressions (which is itself due to the impossibility of an analytical integration of the system’s differential equations). This concept is easily verified for the case of the only reaction whose solution is analytically derived and the expressions of its parameters are known, namely, the primary photoprocess (Maafi and Brown, 2007). Indeed, among all possible sets of parameters of this reaction’s Eq. 4 (which is a mono-Φ-order equation model) that generate good fit of an RK trace, it is possible to single out the true solution set of fitting parameters’ values (ω10λirr, ccλirr, and k10λirr) based on the fact that the equations of the latter coefficients are known (Maafi and Brown, 2007).

In this context, it is important to stress that such an identifiability issue cannot be solved by considering a combination of traces obtained at “any” number of different irradiation wavelengths, as might have been previously proposed (Crano and Guglielmetti, 2003; Delbaere et al., 2011; Micheau et al., 2014). The reason for this inconsistency is a proportional increase of the unknown parameters with each additional Eq. 4 when increasing the number of traces at different λirr. The seemingly success of the aforementioned approach came at a cost: it must assume that the quantum yields of the individual reaction steps are necessarily invariants with λirr (i.e., each species has a unique quantum yield for all λirr, or ΦYjλirr=ΦYj). The latter condition addresses a very particular situation, which, in any case, needs to be proven experimentally and not assumed (the description of the latter approach (Crano and Guglielmetti, 2003) has neither proven experimentally such an invariability of the quantum yield nor considered the necessity of recommending an experimental proof on the quantum yield invariability before envisaging its application).

In a previous study by our team (Maafi and Brown, 2008), detailed spectrokinetic methods have been proposed with the aim of solving the identifiability problem (i.e., leading to the true solutions) of photoreversible systems. They showed the limits of using several traces obtained for a reactive system exposed to different irradiation wavelengths.

For the temporal variation in the total absorbance of the reactive medium (Atotλirr/λobst) irradiated at a given λirr and monitored at λobs, where λirr and λobs may or may not be equal (i.e., εYjλobs and lobs might be different from those measured at λirr), Eq. 4 equivalent takes the following form:

Atotλirr/λobst=Atotλirr/λobs+i=1nΦAωi,AλirrLog1+ccAλirrekiAλirrt.(8)

Note that Eq. (8) has the template of Eq. 4 because the former is a linear combination of the latter for a given reactive system. Hence, the number of terms in Eq. (8) (iA) would, at most, be equal to the number of reaction steps occurring in the photoreaction (nΦ). For the trace of Atotλirr/λobst (Eq. (8)), Atotλirr/λobs is the total absorbance of the reactive medium at the end of the reaction, the factor ωi,Aλirr is equivalent to the ωijλirr of the species individual traces in Eq. 4 multiplied by the respective constants εYjλobslobs, the factor ccAλirr is also constant in all terms of Eq. (8), and the rate constants of the different reaction’s regimes, in Eqs. (4) and (8), are invariant, i.e., kiAλirr=kijλirr.

It is important, however, to mention that the total absorbance trace varies according to the conditions in which the measurement is performed. Atotλirr/λobs is depending on which of the following combinations is considered: i) λirr=λobs and lirr=lobs; ii) λirr=λobs and lirrlobs; iii) λirrλobs and lirr=lobs; iv) λirrλobs and lirrlobs. Hence, the properties of Atotλirr/λobst traces, and consequently, the fitting parameters, except ki,Aλirr, will change between situations (i–iv), for a given reactive system.

Overall, however, the fitting of the RK traces of given systems, with Eq. 4 models, will allow us to describe some general aspects of photokinetics under monochromatic light, as well as to construct a method to fully solve such a photokinetics even when Φλirr is variable with λirr, as it will be discussed in the following sections.

3.6 Metrics in photokinetics

Basically, there are two types of coefficients that fully define the kinetics of a photoreaction: extrinsic and intrinsic coefficients. The former, such as CX0, P0λirr, lirr, and lobs, are controlled by the experimentalist and represent the experimental conditions (assumed known). The latter ones are specific features of the different species involved in the photoreaction, i.e., εYjλirr, ΦYjYjλirr (with jj).

In general, one of the metrics in kinetics is the reaction rate constants (k). However, this typical approach is compromised when the identifiability issue cannot be solved, as previously seen for the fitting of the photokinetic RK traces with Eq. 4. In order to circumvent such a hurdle, the use of the initial reaction rate of the reactant and/or those of the species emanating from the reactant would represent a more appropriate means to quantifying the variability of the reaction speed.

It is to be noted that this metric (e.g., RK:r0Xλirr) is the only accessible trace parameter by numerical integration (the rate constants of the reaction regimes are, in general, not calculated by NIMs and not possibly calculated by the experimentalist from those traces or NIM-fed parameters since their explicit formulae are not known). Therefore, since the initial species’ rates can readily be calculated directly by NIMs (RK:r0λirr) for all reactive species, they become good photokinetic metric tools, indicative of the reaction performance. The proportionality between r0Xλirr and kiXλirr, as indicated by Eq. (6), validates the characterisation of the reactant’s photoreactivity by its initial rate.

Overall, the speed up (or the slowdown) of the reactant photoconversion is expected to be followed by the rest of the reaction species’ reactivities in relative proportions (for instance, a change in the rate when a value of an extrinsic coefficient is changed). Consequently, the variation in the reactant speed in one direction is indicative of the variation in the velocities of each of the subsequent species of the reaction in the same direction.

For the present study, another way is offered for the determination of the initial reaction rate, that is, by applying Eq. 4 to the RK trace of the reactant. The negative initial reactant rate [Eq. (9)], for the Φ-shaped reaction mechanism (Scheme 1) where X is depleted to form Y1 and Y3, can be worked out from the following expression:

Fit:r0Xλirr=Fit:r0Y1λirr+Fit:r0Y3λirr=1ln10ccλirr1+ccλirri=1i0ωi0λirrki0λirr,(9)

where the individual, positive, and initial photoproduct rates (Fit:r0Y1λirr and Fit:r0Y3λirr) are obtained from their respective Eq. 4 as

Fit:r0Y1λirr=1ln10ccλirr1+ccλirri=1i1ωi1λirrki1λirr,(10)
Fit:r0Y3λirr=1ln10ccλirr1+ccλirri=1i3ωi3λirrki3λirr.(11)

It is important to notice that the values of Fit:r0λirr obtained in this way will not vary for the different sets of fitting parameters possibly generated for the RK traces (due to the occurrence of the identifiability issue), as long as each of the sets produces an excellent fit of the RK traces by the corresponding Eq. 4.

An additional way to confirm the validity of the aforementioned equations is to compare the values they generate with those obtained from the theoretical rate law of the reaction [Eq. (1)]. The values of Theo:r0Xλirr, Theo:r0Y1λirr, and Theo:r0Y3λirr can be obtained from the following equation:

r0Xλirr=r0Y1λirrr0Y3λirr=ΦXY1λirr+ΦXY3λirrP0λirr110AXλirr0.(12)

The reliability of our methodology should be confirmed by the equality Fit:r0Yjλirr=Theo:r0Yjλirr=RK:r0Yjλirr in all circumstances (Figure 2).

FIGURE 2
www.frontiersin.org

FIGURE 2. Linear correlation between the values of Theo:r0Yjλirr against both RK:r0Yjλirr and Fit:r0Yjλirr. Data belong to more than 60 reaction cases.

Accordingly, and despite the identifiability problem, it is possible to consistently evaluate the kinetics of a photoreaction, and therefore, r0λirr becomes one of the essential metric tools of photokinetics. In the case of a total absorbance trace, the initial velocity can be obtained from the following expression:

Fit:r0Aλirr=1ln10ccAλirr1+ccAλirri=1iAωiAλirrkiAλirr=Theo:r0Aλirr=lirrorobsj=01,3εYjλirrorobsr0Yjλirr.(13)

It is to be noted that the expressions for the initial rates of the reactant and photoproducts (X, Y1, and Y3) cannot be obtained from the fitting of the total absorbance trace [second term of Eq. (13)] because the theoretical equation of the initial rate is a combination of terms as given by the last term of Eq. (13).

Nonetheless, r0Aλirr is an as good metric for photokinetics as was Fit:r0Xλirr, but with a considerable advantage, that is, of relieving the experimentalist from obtaining the individual species CYjλirrt traces (Atotλirr/λobst can be generated on a routine spectrophotometer, preferably with an in situ irradiation).

3.7 Contribution of spectators to photo-rates

A spectator molecule (one of w molecules, SPMr, with wr1) is a species present in the reactive medium but does not interact with reactants and products of the photoreaction, has a constant concentration, is photo- and thermally inert, but absorbs at the irradiation wavelength. In thermal kinetics, a spectator of this type would have no effect on the reactivity of the species involved in the reaction mechanism. However, it is different in photokinetics. In the cases where a non-zero absorbance of the spectator molecule(s) at λirr, ASPMrλirr0 (Eq. (14)), is considered, it must necessarily make part of the total absorbance of the medium (Eqs. 3, 14), since the fraction of the light absorbed by the spectators is lost for the reactive species. It is to be noted that no conditions on ASPMrλirr to belong to the linearity ranges of the SPMr calibration graph are necessary, since ASPMrλirr does not contribute to the mathematical formalism of the photokinetics other than by its actual constant value.

Atotλirrt=r=1wASPMrλirr+j=0nspAYjλirrt.(14)

For this reason, an increase in the absorbance of the spectators is expected to result in a decrease of all species’ reactivities (i.e., rate constants and initial rates). Figure 3 depicts such a behaviour, where the rate of the reactant X and photoproduct, Y7, are reduced with an increase in the concentration of the spectator for the illustration mechanism shown in Figure 4 (a similar trend is also observed for the rest of the species rates, whose traces are not shown in Figure 3).

FIGURE 3
www.frontiersin.org

FIGURE 3. Evidence of a gradual decrease in X and Y7 reactions’ rates upon an increase of the spectator molecules’ absorbance at the irradiation wavelength, ASPMrλirr. The values of ASPMrλirr are shown. The arrows indicate the directions of evolution of each species traces with increasing values of ASPMrλirr. The data correspond to a multi-consecutive reaction involving four photoproducts, as shown in Figure 4.

FIGURE 4
www.frontiersin.org

FIGURE 4. Sequential percentage reduction of the reactant initial speed with an increase in the absorbance of the spectator molecules (ASPMrλirr) present in the reactive medium. The percentage reduction of r0Xλirr, represented by the dots, was obtained by using RK-calculated and Eq. 4 (RK:r0Xλirr and Fit:r0Xλirr) initial reactant-rate values, and the line corresponds to Theo:r0Xλirr given by Eq. 12.

This expectation has turned out to be true for all the cases investigated. It corroborates the experimental results obtained for the primary photoprocess (XY,εY=0) (Maafi and Maafi, 2013; Maafi and Maafi, 2015a) and the photoreversible reaction (Maafi and Maafi, 2015b), where the effects were quantitatively demonstrated by plotting the rate constant versus the photokinetic factor. However, in the present study, the evidence is provided by a progressive decrease in the initial reactant rate with an increase in SPM concentration.

As shown in Figure 4, each data point corresponds to both RK:r0Xλirr and Fit:r0Xλirr, whereas the curve joining the data points corresponds to Theo:r0Xλirr as functions of ASPMrλirr.

Beyond an established effect of the spectator on photoreactivity, it is important to stress a practical aspect that might be found useful for some applications. In other words, the photoreactivity of a molecule can effectively and significantly be reduced or even virtually stopped by an adequate selection of the concentration of one or more convenient spectator molecules. The reduction in photoreactivity by the presence of a spectator molecule has widely been exploited in chromatic orthogonality by performing a particular reaction within multi-component mixtures, though this use of SPM was treated only through a qualitative approach that has not been mathematically formalised (Bochet, 2006; Hansen et al., 2015).

3.8 CXλirr0, lirr, and εXλirr effects on photoreactivity

The photokinetics of reactions must be dependent on the initial concentration of the starting reactant owing to the rate law [Eq. (1)], and more specifically, its photokinetic factor [Eq. (2)], being a function of concentration. Theoretically, the Φ-order kinetic pattern of the traces should be preserved when changing CXλirr0, but the rate of the reaction should be affected by the value of CXλirr0 (all remaining extrinsic factors being the same), together with the initial reactant rate (as our metric tool in this work). It is also predictable that such an effect of CXλirr0 on the rate of the reactant would be passed on to the photokinetics of the photoproducts whose rijλirr would be proportionally changed according to their particular kinetic properties (since their rate laws depend on the concentrations of all species including that of X). In this configuration, an increase in r0Xλirr is expected for an increased CXλirr0 (i.e., a higher CXλirr0 leads to a higher Atotλirr0=AXλirr0 and, hence, higher values of [110AXλirr0), Eq. (1)]. At the same time, increasing CXλirr0 will cause a reduction in the rate constants (kijλirr) due to smaller values of PKF (Eq. (2), when Atotλirrt increases) as was suggested by the semi-empirical method (Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2014b; Maafi and Maafi, 2015a).

A confirmation of CXλirr0 effects on r0Xλirr was investigated by performing RK calculations. The absolute values of RK:r0Xλirr for a given reaction increase with an increase in the values of CXλirr0, tending asymptotically towards a limit (the latter should be reached at a total absorption of the light). When the RK traces were fitted by the appropriate Eqs 4 or 8, a gradual reduction in kijλirr values was recorded with an increase in the values of CXλirr0 (Figure 5). These trends were independent of the actual mechanism of the photoreaction (an example is provided in Figure 5). Such general results on the effect of CXλirr0 set out a fundamental photokinetic property, which does not seem to have been reported, thus far, in the literature as a general feature, i.e., irrespective of the mechanism of the reaction at hand. The auto-photostabilisation of the reaction with increasing CXλirr0 has previously been experimentally evidenced for two reactions (using semi-empirical equations), namely, the primary photoprocess with an εY1λirr0 (Maafi and Maafi, 2013; Maafi and Maafi, 2014a) and the photoreversible reaction (Maafi and Maafi, 2015a), whose k10λirr linearly correlated with PKF0. Incidentally, it has been observed that the ratios of pairs of species’ final concentrations (end products or photostationary species) were independent of the values of CXλirr0.

FIGURE 5
www.frontiersin.org

FIGURE 5. Variation in the initial reactant rates RK:r0Xλirr and Fit:r0Xλirr, the initial rate of the total absorbance Fit:r0Aλirr, and the fitting rate constant (k10λirr) of the traces, with an increase in reactant initial concentration, for the indicated reaction.

The results obtained for RK:r0Xλirr have served to the endorsement of both Fit:r0Xλirr and Theo:r0Xλirr results (Figure 5), hence confirming the usefulness and reliability of the proposed model equation (Eq. 4).

Furthermore, since lirr is imbedded within the absorption formula [always multiplying the concentration in Eq. (1)], its variation will cause similar photokinetic effects to those described previously for a change in CXλirr0. This was observed by monitoring RK:r0Xλirr and Fit:r0Xλirr, indicating that Eq. 4 applies irrespective of the reactor’s length.

It is to be noted that the effect of varying CXλirr0 and/or lirr is equivalent to when a similar change in the ελirr values occurs, since the former quantities always multiply the latter in Eq. (1) (within the expressions of absorbances). Therefore, investigating a change in CXλirr0 both informs about the trend of varying lirr and equates to studying a class of a reactive system, whose reactant and photoproduct absorption coefficients (ελirr) were proportionally changed. This might be of interest when designing a photoreactive system for particular applications.

3.9 Photokinetic impact of P0λirr, Sirr, and Virr: kinactinometry

For the purpose of describing how the radiation intensity within the reactor, P0λirr, might affect the photoreaction reactivity, it is useful to define what this quantity stands for.

As P0λirr is certainly related to the energy supplied by the light source, the latter is a good starting point.

The incident radiation of the monochromatic light beam, which can be measured by a physical actinometer as a spectral irradiance of the light source, Espirr, is expressed in Wcm2nm1 or Js1cm2nm1 (i.e., energy (J) per unit time, t (s), per unit area [1 cm2), and per wavelength, λirr (nm)]. Since the radiation must be monochromatic for the kinetic model considered in the present work, the measured spectral irradiation corresponds to a single wavelength. This allows to ignore the unit, per wavelength (nm1), in the dimension of the spectral irradiation. The unit J/s corresponds to an energy flux.

In experimental photochemistry, the number of photons is preferred to the energy of the beam since the former allows an adequate quantitative metric because the reaction occurs between this number of photons and the number of photoactive species in the medium. This number of photons is determined as follows.

The energy (Ehνλirr in J) carried by one photon of the radiation is a function of the wavelength λirr.

Ehνλirr=hcλirr,(15)

where h is Planck’s constant expressed in Js (h=6.626081034Js), c is the velocity of light given in m/s (299792458ms1), and λirr is the wavelength in m.

For an einstein (a mole) of photons of a given wavelength (or Avogadro’s number of photons per mole, Na, Na=6.022141023, in einstein1), carrying a total energy Emolhνλirr, the latter equation becomes

Emolhνλirr=Nahcλirr.(16)

Applying this formula to the physically measured energy of the radiation, provided by the spectral irradiance, Espirr, gives the number of photons (number of einstein or moles of photons) carried by the radiation per units of time and area. This quantity represents the incident photon flux of the monochromatic light source through a unit area, flxhνλirr.

flxhνλirr=EspirrEmolhνλirr=EspirrλirrNahc.(17)

The dimension analysis of Eq. (17), flxhνλirr=Espirrλirr/Nahc=Js1cm2m/einstein1Jsms1=einsteins1cm2, confirms a flux dimension (einsteins1) per unit area.

Therefore, in order to express the number of photons entering the reactor per s for a given experimental setup, with the specific irradiated area of the sample, it is necessary to multiply the incident flux per unit surface flxhνλirr, by the actual irradiated area of the sample, Sirr (in cm2).

flxhν,Sirrλirr=flxhνλirrSirr.(18)

Finally, flxhν,Sirrλirr is to be reported to the actual volume (Virr in dm3) of the sample subjected to irradiation. The derived quantity, representing the P0λirr in Eq. (2), is expressed as

P0λirr=flxhνλirrSirrVirr,(19)

whose dimension analysis gives P0λirr=einsteins1cm2cm2/dm3=einsteins1dm3. P0λirr is the number of photons entering the reactor per second, per irradiated area and volume of the investigated sample.

The dimension of P0λirr is adequate for a rate law describing a photochemical reaction [Eq. (1)]. The dimension characterising the left-hand side term of that differential equation, corresponding to a differentiation of the concentration with time, as mol/Ls or moldm3s1, and the right-hand term of the equation has the unit of P0λirr (both quantum yields, Φλirr, and factors multiplying P0λirr [Eq. (2)] are dimensionless). Therefore, the same unit is found for either side of the differential equation (the similarity claim is consistent, since for photons count, mol is equivalent to einstein).

The IUPAC (Braslavsky, 2007) defines a quantity qn,p,λ0 as the radiation spectral photon flux on amount basis, with a dimension of einsteins1nm1 or einsteins1 for a monochromatic light. In the expression given therein for the rate law, the spectral photon flux is divided by the volume irradiated (qn,p,λ0/V, in einsteins1dm3). The latter quantity has the same dimension as P0λirr [Eq. (19)]. However, it is to be noted that Eq. (19) takes into account the actual irradiated area of the sample (the two equations will be equal if Sirr=1cm2).

Therefore, the value of P0λirr can change due to a variation in either the incident flux of the light source or lamp (flxhνλirr), the irradiated area (Sirr), or the volume (Virr) of the sample exposed to the light. In this context, it is obviously important to ensure a vigorous stirring of the reactive medium, throughout irradiation, for consistency with the value of Virr in Eq. (19). It is to be noted also that the ratio Virr/Sirr, in Eq. (19), is not necessarily equal to the optical path length of the light beam inside the sample lirr (it would only be the case if the whole volume is exposed to the light through one of its full sides, but in general, Virr/Sirrlirr).

This, then, commends that the three properties are reported in the investigation in order to complete the adequate photokinetic analysis. Unfortunately, in many published studies, these quantities are missing. This might explain, at least in part, the difficulties in comparing the photokinetic results that were obtained in different laboratories.

Practically, the changes in flxhνλirr, Sirr, and Virr can be controlled by the experimentalist by, respectively, varying the electric potential of the lamp, the blinded surface, or volume irradiated of the sample (these are part of the extrinsic parameters).

The qualitative effect of increasing P0λirr is an increase of the photoreaction reactivity. The quantitative assessment of such an effect is provided by proportional changes in the values of kijλirr (as proven in the previous work (Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2014b; Maafi and Lee, 2015a; Maafi and Maafi, 2015a; Maafi and Lee, 2015b; Maafi and Maafi, 2016; Maafi and Al-Qarni, 2019)). In addition, an increase in P0λirr leads to a reduction in the overall time necessary for the reaction to reach completion and, hence, a reduction in t1/2. Furthermore, a variation in P0λirr causes a proportional change in the value of r0λirr.

The determination of the values of P0λirr is essential for quantitative photochemistry and fundamental in converting a reactive system into an actinometer. The value of P0λirr is obtained from Eq. (20) [which is a rearranged Eq. 12]. Hence, photokinetic data are used to perform actinometry, and this methodology is dubbed here: kinactinometry.

P0λirr=Theo:r0XλirrorFit:r0XλirrΦXY1λirr+ΦXY3λirr110AXλirr0.(20)

Actinometry is the measure of the incident light flux per unit area and volume (P0λirr) at λirr, entering a given reactor (a slab-shaped reactor for our study), and a chemical actinometer is a standardised reactive system that can deliver P0λirr. The standardisation can be achieved kinetically by quantifying the linearly correlated variation of P0λirr with the rate constant of a given reaction step (kijλirr) of the global photoreaction, and/or the initial reactant rate (r0Xλirr or r0Aλirr, the metric in this study). Such a linear correlation is an advantage provided by a description of the photokinetic traces with equations of the type of Eq. 4. The correlation between the number of photons entering the reactor, P0λirr, and r0Yjλirr follows from the relationship between P0λirr and the number of photons absorbed by the considered species, PaYjλirr.

Technically, when varying P0λirr, the model equations (Eq. 4) of the different species of the reaction share the same factors C,jλirr, ωijλirr, and ccλirr since all are independent of P0λirr. The final concentrations (at t=) of the reactive species (C,jλirr) are invariant with P0λirr, but they are reached faster (shorter t1/2) for higher values of P0λirr. Hence, changing P0λirr will induce a proportional change in the kijλirr factors, as the only changing fitting parameter, which adds robustness of the methodology and, overall, eases the fitting process.

Since the aforementioned behaviour is expected for each of r0Yjλirr factors (e.g., for X, Y1, and Y3), then each is a good metric tool for actinometry (with each r0Y0,1or3λirr factor having a different proportionality to P0λirr). A feature that can be beneficial experimentally as it means that a single species’ trace (e.g., the reactant’s) would be sufficient for achieving an actinometric measurement. In addition, it is important to use the total absorbance trace to demonstrate its usefulness for actinometry. It turns out that the fitting parameters, kijλirr factors and r0Aλirr, of the total absorbance traces with increasing P0λirr values also deliver linear correlations (Figure 6). This is an additional confirmation of the validity of the model, but, as importantly, it proves the usefulness of a simplified way to experimentally achieve actinometry. Indeed, all is required is to fit the Atotλirrt trace with Eq. 4. This kinactinometric approach has never been reported in the literature.

FIGURE 6
www.frontiersin.org

FIGURE 6. Linear correlations of the ki0λirr and r0,Xλirr, r0,Y1λirr, r0,Y3λirr, and Fit:r0,Aλirr with various P0λirr values for a reaction governed by the Φ-shaped mechanism shown in Scheme 1. Each point of r0,Yjλirr in the plot cumulates RK, Theo, and Fit values (fitting CXλirrt, CY1λirrt and CY3λirrt traces). The ki0λirr factors were obtained from Eq. 4 corresponding to separate fittings of the individual CXλirrt traces. The r0,Yjλirr factors were calculated from the differentiated species’ equations at t=0 (Eqs 913).

The excellent linear correlation of P0λirr with RK:r0Xλirr and Theo:r0Xλirr stands for a confirmation of that proportionality based on RK calculations. The similar correspondence found with Fit:r0Xλirr and Fit:r0Aλirr is not only another validation of the proposed Eq. 4 but also a simple means to perform actinometry and recruit new actinometers. For a given experiment, the value P0λirr is worked out from the specific equation (e.g., examples given in Figure 6) corresponding to the quantity measured from the fitting equation.

Experimentally, one needs to determine the value of P0λirr. Reference and precise values of P0λirr, at given wavelengths, can also be determined by using the set of reliable actinometers that were previously proposed by our team on the basis of semi-empirical equations (i.e., Eq. 4 type (Maafi and Lee, 2015a; Maafi and Maafi, 2015b; Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2014b, Maafi and Maafi, 2015a, Maafi and Maafi, 2016; Maafi and Al-Qarni,2019)).

The large spectral region coverage by these actinometers (ranging between 250 and 580 nm, Figure 7) is useful for virtually any organic photoactive molecules. Because the implementation of these actinometers is very easy, requiring straightforward kinactinometric methods, they are strong alternatives to the ferrioxalate actinometer whose experimental setup is much more demanding. The list of these actinometers (Figure 7) can be extended by new (cheaper, water soluble, nanocarrier, etc.) candidates using the kinactinometric method proposed in the present work.

FIGURE 7
www.frontiersin.org

FIGURE 7. Spectral ranges for a series of existing efficient actinometers.

3.10 On the reactant’s quantum yield determination

The quantum yield of a reaction step (ΦYjYjλirr) is an intrinsic feature of the reagent Yj for the reaction step YjYj at the irradiation wavelength considered. The monochromatic light is a requirement for the determination of the absolute values of the quantum yield (Braslavsky, 2007). Traditionally, ΦYjYjλirr was held as the most important characteristic of a photoreaction (Crano and Guglielmetti, 2003; Tonnesen, 2004; Pianowski, 2022). However, photokinetic analyses have shown that ΦYjYjλirr is a very important factor in defining the reactivity of a photospecies, nonetheless, in only a partial capacity, as it is one of a set of other elements [Eq. (1)]. The explicit expression of the primary photoprocess rate constant, which has been analytically derived (Maafi and Brown, 2007), indicates that the photochemical quantum yield is, at least, one of the four factors defining the reaction rate constant (kXλirr=ΦXλirrεXλirrP0λirrlirrln 10). A similar conclusion has been reached from the expressions of rate constants of the unimolecular (Maafi and Maafi, 2013), reversible (Maafi and Maafi, 2014a), and multiconsecutive (Maafi and Maafi, 2016) photoreactions (where the kijλirr formulae have been obtained by a semi-empirical method).

In experimental photochemistry, several methods have been proposed for the determination of the quantum yield, albeit commonly based on actinometric measurements combined with the total absorption of the light by the reactive medium. A few examples of photokinetically determined absolute ΦYjYjλirr values have been earlier reported (Crano and Guglielmetti, 2003; Tonnesen, 2004; Pianowski, 2022). The advantage of the photokinetic–actinometric method (or kinactinometric method) is its applicability to relatively low concentrated actinometric solutions (as imposed by the linearity range of the calibration graph, which also means partial absorption of light by the medium) and its ability for the determination of a quantum yield value for each of the individual reaction steps occurring in the overall mechanism (Maafi and Lee, 2015a; Maafi and Lee, 2015b; Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2014b; Maafi and Maafi, 2015a; Maafi and Maafi, 2016; Maafi and Al-Qarni,2019).

Achieving such results (determination of ΦYjYjλirr from the analysis of the traces) through the usage of Eq. 4 might seem, at this stage, impossible because the explicit general formula of the factors kijλirr is unknown. Alternatively, in principle, it would be possible to develop a semi-empirical method for each mechanism in order to work out the absolute values of the different ΦYjYjλirr, but this approach would be time consuming.

It is, however, possible to exploit the model Eq. 4 for the determination of the reactant quantum yield. This is achievable based on Eqs. (21) and (22) (assuming that P0λirr is known).

ΦXY1λirr+ΦXY3λirr=Theo:r0XλirrorFit:r0XλirrP0λirr110AXλirr0.(21)

The individual quantum yields of the divergent reaction steps (Y1XY3) can similarly be worked out [Eq. (22)] from their respective initial velocities (Fit:r0Y1λirr and Fit:r0Y3λirr).

ΦXY1or3λirr=Fit:r0Y1or3λirrP0λirr110AXλirr0.(22)

Figure 8 shows the efficiency of Eqs. (21) and (22) in estimating the original quantum yield values that fed the RK-generated traces of species X, Y1, and Y3. Accordingly, the present photokinetic methodology of determining the quantum yields of the reactant and its immediate photoproducts is relatively easy and might be found beneficial to experimentalists including non-specialists of photokinetics.

FIGURE 8
www.frontiersin.org

FIGURE 8. Excellent correspondence between the calculated quantum yields of X, Y1, and Y3 using Eqs. 21, 22 (based on Fit:r0Xλirr values) and those feeding the RK calculations of these species’ traces.

3.11 Solving for all species’ individual quantum yields

Fully solving photokinetics of a given reaction would necessarily mean that all extrinsic and intrinsic parameter values are unequivocally determined. The extrinsic parameters are easier to obtain with relatively high accuracy. Intrinsic parameters’ values (εYjλirr and ΦYjYjλirr) are most often, all or in large majority, unknown to the investigator, with being relatively stringent to define with precision if at all definable. In fact, no methodology has ever been standardised to solve photokinetics irrespective of the reaction mechanism. There are, however, a few examples of working approaches addressing a handful reaction cases (Crano and Guglielmetti, 2003; Tonnesen, 2004; Pianowski, 2022). Only a fraction of those were based on photokinetic analysis. The approach presented in Section 3.10 is efficient but limited to the value of the quantum yield of X, failing to inform on the values of the rest of the intrinsic parameters of the reactive system. A more efficient approach, based on a semi-empirical method, readily solved the photokinetics of few reactions (Maafi and Maafi, 2013; Maafi and Lee, 2014a; Maafi and Lee, 2014b; Maafi and Lee, 2015a; Maafi and Lee, 2015a; Maafi and Lee, 2015b; Maafi and Lee, 2016).

The formalism presented hereafter claims, for the first time in the photochemistry literature, to provide a general methodology to extract the full set of intrinsic parameters characterising the reactive species of any photoreaction.

The step-by-step procedure is deployed as follows. Its application considers that i) the investigator knows precisely the mechanism operating the photoreaction, ii) the individual CYjλirrt traces corresponding to each species of the reaction mechanism are experimentally collected. In addition, iii) the medium total absorbance Atotλirrt trace is available, and iv) the values of the extrinsic parameters of the reaction (CXλirr0, P0λirr, lirr, lobs, etc.) are all knowns. Conversely to points i) to iv), the procedure considers that the investigator v) has no indications relative to the values of the intrinsic parameters.

In the present study, we will use the RK-calculated traces but, while applying the solving method derived as follows, we will consider unknown intrinsic parameters’ values (that originally served to feed the RK calculation).

The general procedure for solving the kinetics for a given reaction mechanism, which may include one to eight species and nΦ reaction steps (Scheme 1) subjected to a light beam of wavelength λirr, unravels following three stages.

3.11.1 Stage 0: the unknowns

For a reaction involving nsp species (including the reactant), and nΦ reaction steps, the unknows are nsp absorption coefficients (εYjλirr one per species at λirr) and nΦ quantum yields (ΦYjYjλirr one per reaction step). In total, there are nsp+nΦ unknowns to be determined. The values of extrinsic parameters are all supposed known. For simplicity, the trace for the total absorbance should be collected at λobs=λirr and multiplied by (lirr/lobs), such as Atotλirrt=Atotλirr/λirrt=Atotλirr/λobstlirr/lobs, in order to use Atotλirrt in the following equations.

3.11.2 Stage 1: determination of the fitting equations for the traces

(1) Fitting the nsp CYjλirrt and the Atotλirrt traces with adequate Eq. 4: The resulting nsp+1 equations are recorded with the specific values for ωijλirr, ccλirr, and kijλirr for each species Yj and the total absorbance.

(2) The graph showing the overlapping CYjλirrt traces of the nsp species is drawn.

3.11.3 Stage 2: determination of the nsp unknown absorption coefficients

(3) Values of the total absorbance are measured at nΦ selected time intervals (selected from the traces plot constructed in point (2), using its fitting equation set out in point (1).

(4) Using the fitting equations obtained in point (1), the concentrations of each Yj species for nΦ different time intervals selected in point (3) (a total of nΦ×nsp concentration values, nΦnsp1) are worked out.

(5) Values of total absorbances and species concentrations in a set of corresponding nsp equations of the form given by Eq. (3) (chose nsp equations amongst the nΦ×nsp possible). This will deliver nsp linear but linearly independent equations of Atotλirrt.

6) Sole the system of these nsp linear equations for the nsp values of εYjλirr

3.11.4 Stage 3: determination of the nΦ quantum yields of the individual reaction steps

(7) The individual nΦ rate laws (rYjλirrt) for the reaction species are written down according to Eq. (1) (there can be up to nΦ×nsp different equations of rYjλirrt that are written using nΦ×nsp values of the species concentrations (4), total absorbances (5), and the nsp values of the species absorbance coefficients determined in point (6)). nΦ equations of rYjλirrt are selected for the rest of the treatment.

(8) The numerical values of the nΦ rjλirrt selected in point (7) are calculated, for the individual species Yj, as given by Eq. (5), using the values of CYjλirrt obtained in point (4), the total absorbances values measured in point (5), and the absorbance coefficients derived in point (6).

(9) The same data described in point (8) is used to calculate the values of AYjλirrt=εYjλirrlirrCYjλirrt at the nΦ selected time intervals.

(10) The numerical values of PaYjλirrt at the nΦ selected time intervals are calculated using Eq. 2 and the adequate values of CYjλirrt (4), Atotλirrt (5), εYjλirr (6), and AYjλirrt (9).

(11) The aforementioned numerical values of PaYjλirrt (10) and rYjλirrt (8) are introduced in nΦ parametric rate-law equations (7) that include nΦ unknown quantum yields.

(12) This system of nΦ linear but linearly independent equations is solved for the nΦ values of ΦYjλirr

An illustration of the efficiency of the aforementioned method is provided in Table 1 for the photoreversible reaction (whose RK calculation was fed with CXλirr0=1.58105M, P0λirr=1.25105einsteins1dm3, εXλirr=12004M1cm1, εY1λirr=23123M1cm1, ΦXY1λirr=0.062, ΦY1Xλirr=0.034, and lirr=1.65cm, assumed to be equal to lobs, and λirr=λobs). In this particular case, nsp=nΦ=2, which requires two equations of both Atotλirrt and rXλirrt for a complete elucidation of the kinetics (the former used for the extraction of the values of εXandY1λirr and the latter for those of ΦXY1λirr and ΦY1Xλirr).

TABLE 1
www.frontiersin.org

TABLE 1. Data corresponding to the 12 points of the method solving the kinetics of a photoreversible reaction.

Therefore, it is clear that the aforementioned procedure is able to solve for intrinsic parameters of any reaction photokinetics and, hence, definitely settles this, almost a century-old, problem.

3.12 On the generally adopted quantum yield formula

By considering the rate-law equation of a species Yj [as given by Eq. (1)], it is mathematically consistent to perform an integration of the two terms of the equation, independently from one another, when separation of the variables is feasible. Here, each term of the differential equation is integrated with respect to a different variable. For a rate law, this means that the left-hand side of the equation is integrated relative to the concentration whereas the right-hand-side term is integrated relative to time. However, this imposes that the right-hand-side term is a function of time [not absorbances as stated in Eq. (1)]. In other words, the concentrations making part of the absorbances should be replaced by Eq. 4 that are functions of time. Under these conditions, we can write the integral of the rate law in the following general form:

Cjλirr0CjλirrtdC=i=0nΦjΦYjYiλirr0tPaYjλirrtdt+ΦYiYjλirr0tPaYiλirrtdt.(23)

For the reactant, we can write its equation by also making, in the right-hand-side term, the coefficient corresponding to the first reaction step (XY1) visible as

CXλirr0CXλirrtdC=ΦXY1λirr0tPaXλirrtdt+i=1nΦj1ΦXYiλirr0tPaXλirrtdt+ΦYiXλirr0tPaYiλirrtdt.(24)

Hence, we integrate the left-hand side and rearrange Eq. (24) so that it gives the expression for the initial quantum yield.

ΦXY1λirr=CXλirrtCXλirr00tPaXλirrtdt+i=1nΦj1ΦXYiλirr+ΦYiXλirr0tPaYiλirrtdt0tPaXλirrtdt.(25)

Even though Eqs. (24) and (25) are valid (under the condition stated previously), and the integration of the left-hand term of Eq. (24) is straightforward, it is still difficult, in general, to obtain the values of the integrals of the absorbed light by the different species (PaYiλirrt). This stems from complex formulations of the absorbed light terms as functions of time, when introducing the concentration explicit formula (the corresponding Eq. 4). One then obtains functions that do not have known antiderivatives. Therefore, analytical integrations on the right-hand side of Eqs. (24) and (25) are impossible for all photoreactions but one. Indeed, Eq. (25) is analytically integrated only for the primary photoreaction with a transparent photoproduct (εY1λirr=0). For this particular reaction, Eq. (25) becomes

ΦXY1λirr=CXλirrtCXλirr00tPaXλirrtdt.(26)

After introducing the explicit formula of the concentration of X previously analytically derived (Maafi and Brown, 2007) (which has an Eq. 4 formulation), into the expression of the absorbed light, we obtain

PaXλirrt=P0λirr1110AXλirrt=P0λirr111+10AXλirr01ekXλirrt.(27)

Integration of Eq. (27) relative to dt, and introducing the result in the denominator of Eq. (26), we obtain

ΦXY1λirr=CXλirrtCXλirr0P0λirr/kXλirrAXλirr0ln10ln1+10AXλirr01ekXλirrt.(28)

The values of ΦXY1λirr obtained from Eq. 28 and the fitting parameters of the reactant RK trace are quasi-equal to those originally feeding the RK calculation. Another confirmation of Eq. (28) is brought about by Eq.(29), a rearranged integrated rate law of X (Maafi and Brown, 2007), where its parameters take the values that fed the RK calculation. Despite the notable differences in the format between these two equations [Eqs (28) and (29)], the values they generate for ΦXY1λirr are exactly similar.

ΦXY1λirr=1εXλirrP0λirrlirrln10tln10AXλirrt110AXλirr01.(29)

It is to be noted that when the photoproduct of the primary photoprocess absorbs (εY1λirr0), the integrals on the right-hand side of Eq. (24) [and Eq. (25)] become insolvable by a closed-form integration due to missing known antiderivative. This remains true even if the corresponding Eq. 4 is introduced to replace the concentration of the reactant in the integrals. Unfortunately, the same situation will be faced for virtually all remaining photoreactions. However, after incorporating the adequate Eq. 4 in the expression under the integral, the evaluation of PaYjλirr integrations occurring on the right-hand side of Eq. (24) or (25) can eventually be evaluated numerically. Incidentally, the equation proposed by the IUPAC (Braslavsky, 2007) has never been used for the estimation of the quantum yields of a photoreversible reaction [whereas Eq. (24) should apply].

The aforementioned analysis shows that the quantum yield equation proposed in the IUPAC report (Braslavsky, 2007), which is equivalent to Eq. (26), is applicable to a very specific situation, namely, a primary photoprocess with εY1λirr=0, but is not a general equation as might have been suggested in that document. Instead, the only general equation for the reactant is Eq. (24) [or Eq. (25)] for any reaction mechanism. It is also interesting to observe that the definition put forward by Warburg (1917), Warburg (1920), Warburg (1924), Warburg and Rump (1929), and Rubin and Braslavski (2010) cannot be consistent as it defines the quantum yield as derived by dividing the difference of concentrations (CXλirrtCXλirr0) by the total light absorbed by the studied system (i.e., Paλirr=PaYjλirr). Furthermore, it is evident that the formula of the quantum yield proposed by Allmand (1926) and Rubin and Braslavski (2010), which amended that of Warburg, is only valid for the simplest primary photoprocess (where εY1λirr=0), since this is the only case where the ΦYjλirr is evaluated by the difference of concentrations of X at t=0 and t, dividing the amount of light absorbed by that specific species (X) (e.g., Eq. (25) suggests otherwise for X of other reactions). It might then be useful to recommend that the reference IUPAC document (Braslavsky, 2007) is amended with the aforementioned information.

3.13 On the reactant’s quantum yield variability with the irradiation wavelength

The absolute quantum yield value of a reaction step (ΦYjYjλirr) at λirr is a specific feature of species Yj at that irradiation wavelength. The mathematical formalism adopted in photokinetics does not, a priori, impose conditions on the values of ΦYjYjλirr when λirr is changed. The variation in ΦYjYjλirr with irradiation has been acknowledged in the IUPAC report, but no details were provided (Braslavsky, 2007). Our approach may provide a handy way to test whether ΦXYjλirr is variable with λirr by employing the method described in Section 3.10.

Furthermore, for a reactant trace, the ratio of the initial reaction rates (Theo:r0Xλirr or Fit:r0Xλirr) measured at any two different irradiation wavelengths λirr1 and λirr2, is to be equal to the ratio given by Eq. (30) (that is worked out from Eq.(12)), only and only if ΦXY1λirr and/or ΦXY3λirr are invariant with irradiation wavelength (each has the same value for λirr1 and λirr2). Otherwise, Eq. (30) qualitatively proves that ΦXYjλirr is λirr-dependent (even if the actual absolute values of the individual quantum yields are not yet known). It is to be noted that the last term of the right-hand side of Eq. (30) can be computed from available experimental data.

r0Xλirr1r0Xλirr=ΦXY1λirr1+ΦXY3λirr1ΦXY1λirr2+ΦXY3λirr2P0λirr1110AXλirr10P0λirr2110AXλirr20=P0λirr1P0λirr2110AXλirr10110AXλirr20.(30)

The simplicity of the method is an advantage to quickly test whether the quantum yield is constant with wavelength. This might be of interest to the community as there are not many evaluations of the invariability of the quantum yield with irradiation, probably considered irrelevant owing to the general assumption that such an invariability should be taken for granted. The argument behind such a state of the matter relates to the Kasha–Vavilov rule (Kasha, 1950), despite that this rule concerned fundamentally photophysical processes and has not been expanded, by the authors, to photochemistry. The literature has reported a large number of systems where the quantum yield is not constant over two or more wavelengths (Becker et al., 1969; Becker and Favaro, 2011; Reinfelds et al., 2019; Montalti et al., 2020), including the ferrioxalate actinometer (Montalti et al., 2020), since as early as 1958 (Zimmerman, 1958). In our team, using the semi-empirical method and by systematic screenings, we have observed the variability of quantum yields with λirr for species involved in various mechanisms, whose reactants belonged to several chemical families, and their reactive systems have diverse applications. Linear, triangular, and sigmoid variations of ΦXY1λirr with λirr have been observed (Maafi and Maafi, 2013; Maafi and Maafi, 2014a; Maafi and Maafi, 2014b; Maafi and Lee, 2015a; Maafi and Maafi, 2015a; Maafi and Lee, 2015b; Lee, 2016; Maafi and Maafi, 2016; Maafi and Al-Qarni, 2018; Maafi and Al-Qarni, 2019). It is then reasonable to consider the aforementioned results as a recurrent and objective experimental observation, which cannot be simply dismissed by invoking experimental discrepancies. In addition, up to date, no fundamental explanation has been proposed and accepted by the community for a supposed invariability of Φ with λirr. The approach described above, might be able to help settle the debate in an effective and easy-to-implement way.

The methods and procedures presented in the previous sections prove a number of reaction behaviours and features and describe the ways to quantify them. However, one needs to keep in mind several important considerations. The approaches are built on the consideration that the concentrations of the reactive species of the investigated photoreaction, at any reaction time, all fall within the respective linearity ranges of their individual calibration graphs. Hence, it is highly recommended to experimentally apply these methods only to a system whose reactant initial concentration is relatively low (the lower, the better but, perhaps, where the total absorbance of the medium is, at least, below 0.5 at any reaction time. At such a value of the total absorbance, it might reasonably be assumed that, for a wide range of organic molecules, the concentrations of the species fall within their linearity ranges).

Knowing that the fitting performance of the experimental traces with Eq. 4 is tributary to both the experimental data quality and quantity, it is recommended to ensure such data have high precision.

4 Concluding remarks

Numerical integration (RK-NIM) has been used in the present work not for kinetic elucidation, as usually proposed in the literature, but for the purpose of describing the behaviours of photoreactions in different situations. Based on the initial velocity, as a metric, several reactivity features have been quantified and proven to generally occur for photoreactions, regardless of the governing mechanism.

A model equation, of the Φ-order character, has proven to faithfully reproduce the kinetic traces generated by RK-NIM. It stands for a unifying model able to describe photokinetics in the many reaction conditions and properties laid out here but expected also to be, in general, valid for other situations not described in the present work.

This model equation is a meaningful tool, as it facilitates full solving for the intrinsic parameters of photoreactions, quantifies the effects of various factors influencing reactivity, and delivers kinactinometers, in relatively handy procedures.

Overall, the findings of this work contribute to standardising photokinetic investigation and lay solid grounds for further developments in this important subject.

The application of a similar approach, to that developed here, is ongoing for the photokinetics under polychromatic light.

Furthermore, it is conjectured that the strategy presented here might be of interest to studying experimental setups of different reactor geometry and spatial distribution of the incidence radiation. The Φ-order kinetic character of photoreactions is expected to be preserved under those conditions, for which Eq. 4 template should apply.

Data availability statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Author contributions

MM: conception, design and writing, and approval of the study.

Funding

This work received financial support from De Montfort University.

Conflict of interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Allmand, A. J. (1926). Part I.—einstein's law of photochemical equivalence. Introductory address to Part I. Trans. Faraday Soc. 21, 438–452. doi:10.1039/TF9262100438

CrossRef Full Text | Google Scholar

Becker, R. S., Dolan, E., and Balke, E. B. (1969). Vibronic effects in photochemistry—competition between internal conversion and photochemistry. J. Chem. Phys. 50, 239–245. doi:10.1063/1.1670783

CrossRef Full Text | Google Scholar

Becker, R. S., and Favaro, G. (2011). New concepts in photochemistry and photophysics: Photochromic and other type molecules. J. Photochem. Photobiol. C Chem. Photochem. Rev. 12, 167–176. doi:10.1016/j.jphotochemrev.2011.08.003

CrossRef Full Text | Google Scholar

Bochet, C. G. (2006). Photochemical release of various functional groups. Pure Appl. Chem. 78, 241–247. doi:10.1351/pac200678020241

CrossRef Full Text | Google Scholar

Bonacchi, S., El Garah, M., Ciesielski, A., Herder, M., Conti, S., Cecchini, M., et al. (2015). Surface-induced selection during in situ photoswitching at the solid/liquid interface. Angew. Chem. Int. Ed. Engl. 54, 4865–4869. doi:10.1002/anie.201412215

PubMed Abstract | CrossRef Full Text | Google Scholar

Bonfield, H. E., Knauber, T., Lévesque, F., Moschetta, E. G., Susanne, F., and Edwards, L. J. (2020). Photons as a 21st century reagent. Nat. Commun. 11, 804. doi:10.1038/s41467-019-13988-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Braslavsky, S. E. (2007). Glossary of terms used in photochemistry. IUPAC, Pure Appl. Chem. 79, 293–465. doi:10.1351/pac200779030293

CrossRef Full Text | Google Scholar

Cahova, H., and Jaschke, A. (2013). Nucleoside-based diarylethene photoswitches and their facile incorporation into photoswitchable DNA. Angew. Chem. Int. Ed. 52, 3186–3190. doi:10.1002/anie.201209943

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, X., Wehle, S., Kuzmanovic, N., Merget, B., Holzgrabe, H., Konig, B., et al. (2014). Acetylcholinesterase inhibitors with photoswitchable inhibition of β-amyloid aggregation. ACS Chem. Neurosci. 5, 377–389. doi:10.1021/cn500016p

PubMed Abstract | CrossRef Full Text | Google Scholar

Crano, J. C., and Guglielmetti, R. J. (2003). Organic photochromic and thermochromic compounds. New York: Springer. doi:10.1007/b114211

CrossRef Full Text | Google Scholar

Delbaere, S., Vermeersh, G., and Micheau, J. C. (2011). Quantitative analysis of the dynamic behaviour of photochromic systems. J. Photochem. Photobiol. C Chem. Photochem. Rev. 12, 74–105. doi:10.1016/j.jphotochemrev.2011.05.004

CrossRef Full Text | Google Scholar

Frigoli, M., Jousselin-Oba, T., Mamada, M., Marrot, J., Zangarelli, A., Pannacci, D., et al. (2020). Synthesis and photochromic behaviour of a series of benzopyrans bearing an N-phenyl-carbazole moiety: Photochromism control by the steric effect. Photochem. Photobiol. Sci. 19, 1344–1355. doi:10.1039/D0PP00202J

PubMed Abstract | CrossRef Full Text | Google Scholar

Hansen, M. J., Velema, W. A., Lerch, M. M., Szymański, W., and Feringa, B. L. (2015). Wavelength-selective cleavage of photoprotecting groups: Strategies and applications in dynamic systems. Chem. Soc. Rev. 44, 3358–3377. doi:10.1039/C5CS00118H

PubMed Abstract | CrossRef Full Text | Google Scholar

Hattersley, J. G., Pérez-Velázquez, J., Chappell, M. J., Bearup, D., Roper, D., Dowson, C., et al. (2011). Indistinguishability and identifiability of kinetic models for the MurC reaction in peptidoglycan biosynthesis. Comput. methods programs biomedecine 104, 70–80. doi:10.1016/j.cmpb.2010.07.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Irie, M. (2000). Diarylethenes for memories and switches. Chem. Rev. 100, 1685–1716. doi:10.1021/cr980069d

PubMed Abstract | CrossRef Full Text | Google Scholar

Irie, M. (2015). Discovery and development of photochromic diarylethenes. Pure Appl. Chem. 87, 617–626. doi:10.1515/pac-2015-0208

CrossRef Full Text | Google Scholar

Irie, M., Fulcaminato, T., Matsuda, K., and Kobatake, S. (2014). Photochromism of diarylethene molecules and crystals: Memories, switches, and actuators. Chem. Rev. 114, 12174–12277. doi:10.1021/cr500249p

PubMed Abstract | CrossRef Full Text | Google Scholar

Kasha, M. (1950). Characterization of electronic transitions in complex molecules. Discuss. Faraday Soc. 9, 14–19. doi:10.1039/DF9500900014

CrossRef Full Text | Google Scholar

Kochman, M. A., Gryber, T., Durbeej, B., and Kubas, A. (2022). Simulation and analysis of the relaxation dynamics of a photochromic furylfulgide. Phys. Chem. Chem. Phys. 24, 18103–18118. doi:10.1039/d2cp02143a

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, L. Y. (2016). Study of the photodegradation and photostability of anti-cancer drugs in different media towards the development of both new actinometers and liquid formulations. PhD Thesis. Leicester, England: De Montfort University.

Google Scholar

Maafi, M., and Al-Qarni, M. A. (2019). Dependence of pinosylvin photochemical quantum yield on excitation wavelength. Trends photochem. Photobiol. 18, 45–58.

Google Scholar

Maafi, M., and Al-Qarni, M. A. (2022). Mono- and polychromatic light diarylethene-actinometer for the visible range. Dyes pigments 198, 109942. doi:10.1016/j.dyepig.2021.109942

CrossRef Full Text | Google Scholar

Maafi, M., and Al-Qarni, M. A. (2018). Φ-order spectrophotokinetic characterisation and quantification of trans-cis oxyresveratrol reactivity, photodegradation and actinometry. Spectrochim. Acta A Mol. Biomed. Spectrosc. 188, 64–71. doi:10.1016/j.saa.2017.06.067

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Brown, R. G. (2005b). General analytical solution for the kinetics of AB(k,Φ) and ABC(k,Φ) systems. Int. J. Chem. Kinet. 37, 162–174. doi:10.1002/kin.20061

CrossRef Full Text | Google Scholar

Maafi, M., and Brown, R. G. (2008). Kinetic analysis and elucidation options for AB(1k,2ϕ) systems. new spectrokinetic methods for photochromes. new spectrokinetic methods photochromes. Photochem. Photobiol. Sci. 7, 1360–1372. doi:10.1039/b807556e

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Brown, R. G. (2005a). Photophysics and kinetics of naphthopyran derivatives, part 2: Analysis of diarylnaphthopyran kinetics. Degeneracy of the kinetic solution. Degeneracy Kinet. Solut. Int. J. Chem. Kinet. 37, 717–727. doi:10.1002/kin.20130

CrossRef Full Text | Google Scholar

Maafi, M., and Brown, R. G. (2007). The kinetic model for AB(1Φ) systems: A closed-form integration of the differential equation with a variable photokinetic factor. J. Photochem. Photobiol. Chem. 187, 319–324. doi:10.1016/j.jphotochem.2006.10.030

CrossRef Full Text | Google Scholar

Maafi, M., and Lee, L. Y. (2015b). Actinometric and Φ-order photodegradation properties of anti-cancer Sunitinib. J. Pharm. Biomed. Anal. 110, 34–41. doi:10.1016/j.jpba.2015.02.026

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Lee, L. Y. (2015a). Determination of dacarbazine Φ-order photokinetics, quantum yields, and potential for actinometry. J. Pharm. Sci. 104, 3501–3509. doi:10.1002/jps.24568

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Maafi, W. (2016). Modeling and elucidation of the kinetics of multiple consecutive photoreactions AB 4 (4Φ) with Φ-order kinetics Application to the photodegradation of RiboflavinΦ with Φ-order kinetics. J. Pharm. Sci. 105, 3537–3548. doi:10.1016/j.xphs.2016.06.030

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Maafi, W. (2014b). Montelukast photodegradation: Elucidation of Ф-order kinetics, determination of quantum yields and application to actinometry. Int. J. Pharm. 471, 544–552. doi:10.1016/j.ijpharm.2014.05.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Maafi, W. (2015b). Quantification of unimolecular photoreaction kinetics: Determination of quantum yields and development of actinometers—the photodegradation case of cardiovascular drug nisoldipine. J. Photoenergy 2015, 1–12. doi:10.1155/2015/454895

CrossRef Full Text | Google Scholar

Maafi, M., and Maafi, W. (2015a). Quantitative assessment of photostability and photostabilisation of Fluvoxamine and its design for actinometry. Photochem. Photobiol. Sci. 14, 982–994. doi:10.1039/c5pp00022j

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, M., and Maafi, W. (2014a). Ф-order kinetics of photoreversible-drug reactions. Int. J. Pharm. 471, 536–543. doi:10.1016/j.ijpharm.2014.05.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Maafi, W., and Maafi, M. (2013). Modelling nifedipine photodegradation, photostability and actinometric properties. Int. J. Pharm. 456, 153–164. doi:10.1016/j.ijpharm.2013.07.075

PubMed Abstract | CrossRef Full Text | Google Scholar

Malato, S., Giménez, J., Oller, I., Agüera, A., and Sánchez Pérez, J. A. (2020). “Removal and degradation of pharmaceutically active compounds (PhACs) in wastewaters by solar advanced oxidation processes,” in Removal and degradation of pharmaceutically active compounds in wastewater treatment. The handbook of environmental chemistry. Editors S. Rodriguez-Mozaz, P. Blánquez Cano, and M. Sarrà Adroguer (Cham: Springer). doi:10.1007/698_2020_688

CrossRef Full Text | Google Scholar

H. Mauser, G. Gauglitz, R. G. Compton, and G. Hancock (1998). Comprehensive chemical kinetics, photokinetics: Theoretical fundamentals and applications. (Amsterdam: Elsevier).

Google Scholar

Micheau, J. C., Coudret, C., Kobeleva, O. I., Barachevsky, V. A., Yarovenko, V. N., Ivanov, S. N., et al. (2014). Quantitative study of photochromic transformations of diarylethene derivatives with either perhydrocyclopentene or oxazolone or lactone units. Dyes Pigments 106, 32–38. doi:10.1016/j.dyepig.2014.02.013

CrossRef Full Text | Google Scholar

Montalti, M., Credi, A., Prodi, L., and Gandolfi, M. T. (2020). Handbook of photochemistry. Boca Raton, Florida: CRC Press, 978–0367577902.

Google Scholar

Nakatani, k., Piard, J., Yu, P., and Métivier, R. (2016). Photochromic materials: Preparation, properties and applications. (Hoboken, New Jersey: Wiley).

Google Scholar

Okuda, J. Y., Tanaka, Y., Kodama, R., Sumaru, K., Morishita, K., Kanamori, T., et al. (2015). Photoinduced cytotoxicity of a photochromic diarylethene via caspase cascade activation. Chem. Commun. 51, 10957–10960. doi:10.1039/C5CC02200B

PubMed Abstract | CrossRef Full Text | Google Scholar

Ovchinnikov, A., Pogudin, G., and Thompson, P. (2023). Input-output equations and identifiability of linear ODE models. IEEE Trans. Autom. Contr 68, 812–824. doi:10.1109/TAC.2022.3145571

CrossRef Full Text | Google Scholar

Pianowski, Z. L. (2022). Molecular photoswitches: Chemistry, properties, and applications. Hoboken, New Jersey: Wiley. doi:10.1002/9783527827626

CrossRef Full Text | Google Scholar

Reinfelds, M., Hermanns, V., Halbritter, T., Wachtveitl, J., Braun, M., Slanina, T., et al. (2019). A robust, broadly absorbing fulgide derivative as a universal chemical actinometer for the UV to NIR region. ChemPhotoChem 6, 441–449. doi:10.1002/cptc.201900010

CrossRef Full Text | Google Scholar

Rubin, M. B., and Braslavski, S. E. (2010). Quantum yield: The term and the symbol. A historical search. A Hist. search. photochem. Photobiol. Sci. 9, 670–674. doi:10.1039/b9pp00182d

PubMed Abstract | CrossRef Full Text | Google Scholar

Sarter, C., Heimes, M., and Jaschke, A. (2016). The role of alkyl substituents in deazaadenine-based diarylethene photoswitches. Beilstein J. Org. Chem. 12, 1103–1110. doi:10.3762/bjoc.12.106

PubMed Abstract | CrossRef Full Text | Google Scholar

Scaiano, J. C. (2020). Handbook of organic photochemistry. Boca Raton, Florida: CRC Press.

Google Scholar

Shit, S. C., Shown, I., Paul, R., Chen, K. H., Mondal, J., and Chen, L. C. (2020). Integrated nano-architectured photocatalysts for photochemical CO2 reduction. Nanoscale 12, 23301–23332. doi:10.1039/D0NR05884J

PubMed Abstract | CrossRef Full Text | Google Scholar

Singer, M., and Jaschke, A. (2010). Reversibly photoswitchable nucleosides: Synthesis and photochromic properties of diarylethene-functionalized 7-deazaadenosine derivatives. J. Am. Chem. Soc. 132, 8372–8377. doi:10.1021/ja1024782

PubMed Abstract | CrossRef Full Text | Google Scholar

Singer, M., Nierth, A., and Jaschke, A. (2013). Photochromism of diarylethene-functionalized 7-deazaguanosines. Eur. J. Org. Chem. 2013, 2766–2769. doi:10.1002/ejoc.201300261

CrossRef Full Text | Google Scholar

Tonnesen, H. H. (2004). Photostability of drugs and drug formulations. Boca Raton: CRC Press.

Google Scholar

Vajda, S., and Rabitz, H. (1988). Identifiability and distinguishability of first-order reaction systems. J. Phys. Chem. 92, 701–707. doi:10.1021/j100314a024

CrossRef Full Text | Google Scholar

Velema, W. A., Szmanski, W., and Ferringa, B. (2014). Photopharmacology: Beyond proof of principle. J. Am. Chem. Soc. 136, 2178–2191. doi:10.1021/ja413063e

PubMed Abstract | CrossRef Full Text | Google Scholar

Warburg, E. (1924). Die quantenregeln in der Photochemie. Naturwissenschaften. 12, 1058–1063. doi:10.1007/bf01452119

CrossRef Full Text | Google Scholar

Warburg, E. (1920). Electrochem 54–59.

CrossRef Full Text | Google Scholar

Warburg, E. (1917). Über die Anwendung der Quantenhypothese auf die Photochemie. Die Naturwiss 5, 489–494. doi:10.1007/BF02450059

CrossRef Full Text | Google Scholar

Warburg, E., and Rump, W. Z. (1929). Über die Photolyse der Lösungen von Schwefelwasserstoff in Hexan und in Wasser und über die Photolyse von Lösungen im allgemeinen. Z. für Phys. 58, 291–305. doi:10.1007/bf01340381

CrossRef Full Text | Google Scholar

Weerasekara, R. K., Uekusa, H., and Hettiarachchi, C. V. (2017). Multicolor photochromism of fulgide mixed crystals with enhanced fatigue resistance. Cryst. Growth Des. 17, 3040–3047. doi:10.1021/acs.cgd.6b01708

CrossRef Full Text | Google Scholar

Ziani, Z., Loiseau, F., Lognon, E., Boggio-Pasqua, M., Philouze, M. C., Cobo, S., et al. (2021). Synthesis of a negative photochrome with high switching quantum yields and capable of singlet-oxygen production and storage. Chem. Eur. J. 27, 16642–16653. doi:10.1002/chem.202103003

PubMed Abstract | CrossRef Full Text | Google Scholar

Zimmerman, G., Chow, L. Y., and Paik, U. J. J. (1958). The photo-chemical isomerization of azobenzene. J. Am. Chem. Soc. 80, 3528–3531. doi:10.1021/ja01547a010

CrossRef Full Text | Google Scholar

Glossary

Keywords: Φ-order kinetics, photokinetics, quantum yield, actinometry, solving kinetics, monochromatic light

Citation: Maafi M (2023) On photokinetics under monochromatic light. Front. Chem. 11:1233151. doi: 10.3389/fchem.2023.1233151

Received: 01 June 2023; Accepted: 22 August 2023;
Published: 14 September 2023.

Edited by:

Yuanhai Su, Shanghai Jiao Tong University, China

Reviewed by:

Fang Zhao, East China University of Science and Technology, China
Chunhui Dai, East China University of Technology, China

Copyright © 2023 Maafi. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Mounir Maafi, mmaafi@dmu.ac.uk

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.