Your new experience awaits. Try the new design now and help us make it even better

ORIGINAL RESEARCH article

Front. Astron. Space Sci., 30 July 2025

Sec. Astrochemistry

Volume 12 - 2025 | https://doi.org/10.3389/fspas.2025.1570525

This article is part of the Research TopicPhosphorus Chemistry in the Interstellar Medium and Planetary AtmospheresView all 5 articles

Strong parameter hierarchy in the interstellar phosphorus chemical network

  • 1Centro de Astrobiología (CAB), CSIC-INTA, Madrid, Spain
  • 2Grupo Interdisciplinar de Sistemas Complejos (GISC), Madrid, Spain
  • 3Instituto de Investigación Tecnológica (IIT), Universidad Pontificia Comillas, Madrid, Spain

Phosphorus-bearing molecules are fundamental for life on Earth, yet their astrochemical origins remain poorly understood. Their formation in the interstellar medium has been challenging to elucidate due to limited spectroscopic detections and the reliance on theoretical models that depend on numerous kinetic parameters whose values are very uncertain. Multi-parameter models often suffer from “sloppiness”, where many parameter combinations exhibit negligible influence on model outcomes, while a few dominate system behavior. In this study, we introduce the Fisher Information Spectral Reduction (FISR) algorithm, a novel and computationally efficient method to reduce the complexity of such sloppy models. Our approach exposes the strong parameter hierarchy governing these systems by identifying and eliminating parameters associated with insensitive directions in the parameter space. Applying this methodology to the phosphorus astrochemistry network, we reduce the number of reaction rate coefficients from 14 to 3, pinpointing the key reactions and kinetic parameters responsible for forming PO and PN, the main phosphorus-bearing molecules typically detected in interstellar space. The simplified model retains its predictive accuracy, offering deeper insights into the mechanisms driving phosphorus chemistry in the interstellar medium. This methodology is applicable to multi-parameter models of any kind and, specifically in astrochemistry, facilitates the development of simpler, more realistic and interpretable models to effectively guide targeted observational efforts.

1 Introduction

Phosphorus (P) is an element of significant astrobiological importance due to its abundance in biomass and its critical role in essential biochemical functions: the backbone of nucleic acids contains sugar-phosphates, phosphorylated molecules act as energy carriers, and cellular membranes contain phospholipid head groups (Walton et al., 2023). Interestingly, despite its ubiquity in life on Earth, phosphorus is far less abundant on cosmic scales than other essential elements for life, such as hydrogen (H), carbon (C), oxygen (O), and nitrogen (N) —a phenomenon often referred to as the phosphorus enigma. In fact, phosphorus ranks 18th in cosmic abundance, orders of magnitude lower than other biogenic elements (Maciá-Barber, 2020). With regard to the origin of terrestrial phosphorus, it is worth noting that the chemical complexity of early Earth was augmented by the influx of molecular compounds inherited from its parental molecular cloud during the process of planetary formation and the subsequent impact of asteroids and comets on its surface. Interplanetary dust, meteorites, and large impactors may have deposited reactive, reduced oxidation state phosphorus onto early Earth’s surface. This process, supplemented by terrestrial P-reduction pathways, enriched prebiotic environments with reactive P, including organic phosphonates with carbon–phosphorus bonds, as identified in the Murchison meteorite (Cooper et al., 1992). Furthermore, the recent detection of PO in comet 67P/Churyumov–Gerasimenko (Rivilla et al., 2020) and of phosphorus-rich grains in asteroid Ryugu samples (Pilorget et al., 2024), indicates that minor Solar System bodies can act as reservoirs for phosphorus-bearing compounds. Therefore, the detection of phosphorus sources in the interstellar medium (ISM) is key to trace the evolution of phosphorus during the Solar System formation and unveil how this element became available for the origin of life.

In recent decades, the interstellar chemistry of phosphorus has emerged as a promising area of research driven by high-sensitivity instrumentation such as the Atacama Large Millimeter/submillimeter Array (ALMA). To date, more than 300 molecules have been detected in interstellar and planetary environments, including complex organic molecules (COMs) with potential prebiotic significance. Despite this impressive detection capacity, only seven P-bearing molecules have been unambiguously identified in the ISM: PO (Tenenbaum et al., 2007; Rivilla et al., 2016), PO+ (Rivilla et al., 2022), PN (Turner and Bally, 1987; Ziurys, 1987; Fontani et al., 2016), CP (Guelin et al., 1990), HCP (Agúndez et al., 2007), CCP (Halfen et al., 2008), and PH3 (Agúndez et al., 2008; Tenenbaum and Ziurys, 2008). Among these, only PO, PO+, and PN have been detected in star-forming regions, but with low abundances (the detected abundances of PO and PN relative to H are on the order of 1010 and 1011, respectively; see e.g., Lefloch et al. (2016), Rivilla et al. (2018), Rivilla et al. (2020), Bernal et al. (2021), Fontani et al. (2024), Lefloch et al. (2024), Scibelli et al. (2025)). In addition, astronomical observations show that PO is systematically more abundant than PN, with abundance ratios of [PO]/[PN]1.43 across various sources (Lefloch et al., 2016; Rivilla et al., 2018; Bernal et al., 2021).

Besides the discovery of new chemical species in the ISM through spectroscopic techniques, astrochemistry also relies on laboratory experimental work and computational models. Laboratory experiments replicate interstellar physical conditions to obtain rotational spectra for known molecular compositions. Computational models, on the other hand, integrate data from astronomical measurements and laboratory experiments, providing a framework to hypothesize the underlying physical-chemical processes responsible for the observed molecular abundances. In this context, astrochemical models such as UCLCHEM (Holdship et al., 2017) or Nautilus (Ruaud et al., 2016) have been developed to simulate the time evolution of abundances for numerous interstellar species, which interact through extensive networks of chemical reactions under various physical conditions and energetic physical processes (e.g., UV/cosmic rays irradiation, protostellar heating or stellar feedback in the form of high-velocity winds inducing shock waves) in both the gas phase and on dust grain surfaces. These models typically consist of applying the law of mass action to all reactions in the chemical network and solving numerically the associated systems of ordinary differential equations. However, the precision of mass-action kinetics depends heavily on the accuracy of the reaction rate coefficients, which remain poorly determined for the majority of reactions.

In the past decade, it has become clear that astrochemical models present severe limitations in reproducing the [PO]/[PN] ratios observed across the ISM (predicted ratios <1 versus the observed ratios of 1.4–3; see e.g., Aota and Aikawa (2012), Jiménez-Serra et al. (2018), Chantzos et al. (2020), Sil et al. (2021)). To get insight into this problem, in a previous work, we conducted a detailed analysis of the chemistry of phosphorus in molecular clouds (Fernández-Ruz et al., 2023). This analysis employed a theoretical approach integrating complex networks, sensitivity analysis, and Bayesian statistics to model 14 key chemical reactions. Unlike traditional numerical approaches in astrochemical modeling, our theoretical method provided deep insights into the role of each reaction in the formation of PO and PN, resolving the discrepancies between modeled and observed abundance ratios. Additionally, this study revealed a strong hierarchical structure in the kinetic parameters of the system, where the formation of PO and PN in molecular clouds was insensitive to a large number of parameter combinations but highly dependent on a selected few. This intrinsic parameter insensitivity aligns very well with the concept of a “sloppy model”, where system behavior is governed by a few “stiff” parameter combinations, while others, referred to as “sloppy directions”, exhibit extreme insensitivity to large-scale fluctuations (Brown and Sethna, 2003; Quinn et al., 2023).

Sloppy models are prime candidates for model reduction techniques. In general, these techniques aim to simplify large-scale dynamical systems through the elimination of unnecessary parameters while preserving essential characteristics (Antoulas, 2005). The primary techniques can be categorized into (i) eigenvector projection methods (using Singular Value Decomposition) and (ii) moment matching methods. Some eigenvector projection methods diagonalize the system and truncate those states that are difficult to control and observe (Moore, 1981), while others aim to minimize an error measure of the system (Glover, 1984) or separate fast and slow timescales (Gugercin and Antoulas, 2004). On the contrary, moment-based methods focus on reducing the complexity while approximating accurately some statistics (often central moments estimated from the data). Overall, both categories of methods strive to achieve a balance between computational efficiency and accuracy, making them suitable for large-scale models but relying too strongly on linearization or on global measurements. Alternatively, M.K. Transtrum and collaborators, following a different approach, developed a decade ago the Manifold Boundary Approximation Method (MBAM) (Transtrum et al., 2011; Transtrum and Qiu, 2014; Transtrum and Qiu, 2016; Quinn et al., 2023). This elegant algorithm is based on information geometrical arguments, balancing the data with the complexity of the model exploiting the low sensitivity to some combinations of parameters.

In this paper, we introduce a methodology for reducing and simplifying models while retaining their predictive accuracy inspired by the MBAM reduction technique and apply it specifically to the phosphorus astrochemistry network. The proposed method, the Fisher Information Spectral Reduction method (hereafter, FISR method), is conceptually simpler than the MBAM and thus can be implemented by an easier and more efficient algorithm. It performs iterative dimensional reduction steps, progressively decreasing the number of model parameters. The final outcome is a reduced model with significantly fewer parameters, in which unnecessary complexity is removed.

Applying the FISR algorithm to the 14-reaction model of phosphorus chemistry in molecular clouds presented in Fernández-Ruz et al. (2023), we demonstrate that the observed abundances of PO and PN can be explained by a much simpler model comprising just 3 reactions and 3 parameters. This simplified model not only identifies the key reactions governing PO and PN formation but also offers a deeper comprehension of the hierarchical structure of the parameter space. Therefore, the goal of this work is not to obtain a model that improves the predictions of existing astrochemical models, but rather to derive simpler and interpretable models that maintain comparable predictive accuracy while eliminating unnecessary complexities. Our findings highlight that the complexity of a model should correspond to the complexity of the available knowledge with which it is constructed. This principle—also known in the sciences as Occam’s razor—, is broadly applicable beyond astrochemistry to other disciplines reliant on modeling.

The organization of this paper is as follows. In Section 2 we present the mathematical foundation of the complexity reduction method introduced here, the FISR method, and outline the algorithmic steps for its general implementation. In Section 3 we apply the FISR method to the phosphorus chemistry in the interstellar medium, uncovering the underlying chemical network dynamics that governs the formation of PO and PN across short and long timescales. In Section 4, we discuss the implications of the main findings for future phosphorus astrochemical research and examine the strengths and limitations of the FISR method within the context of complexity reduction methods.

2 Methods: the FISR algorithm

The method presented in this work is inspired by the Manifold Boundary Approximation Method (MBAM), originally developed and published by Transtrum and Qiu (2014), and shares its basic principles and primary objective. The reasoning behind this method is that models often do not respond significantly to changes in certain parameters. By applying concepts from information geometry, this approach identifies specific trajectories across the statistical manifold that represent meaningful combinations of the parameters. These trajectories follow the so-called sloppy directions, which are directions in the parameter space that are not informative and along which large parameter changes produce only a minimal impact on the model output. The term boundary in the name of the method connects with the practical observation made by the authors that, except in rare pathological cases, the topology of an N-dimensional manifold contains boundaries of N1 dimensions. This is the reason why one can drop one combination of parameters safely when a trajectory over the manifold reaches one boundary, while preserving the model’s predictive power (Transtrum and Qiu, 2014).

Our proposed algorithm, the Fisher Information Spectral Reduction (FISR) method, successfully achieves the same goals as MBAM but differs significantly from it in its technical details and algorithmic procedures. We name it FISR method because it is based on the eigenspectrum of the Fisher Information Matrix (FIM), which quantifies the amount of information a dataset carries about the model parameters, but also provides the directions on the statistical manifold carrying less information (the sloppiest directions). Since high information content is associated with strong parameter influence, the eigenspectrum of the FIM encodes the sensitivity of model outputs to changes in parameter values. In this Section, we present the mathematical framework of the FISR method and outline its algorithmic implementation from a general perspective, making it applicable to any sloppy model.

2.1 Theoretical framework and definitions

Let us consider a model y=y(θ) with a set of N parameters θ=θ1,,θμ,,θN (see Table 1 for a summary of all the symbols used in this Section). The M outputs of the model are y=y1,,ym,yM, where m typically stands for time values of the evolution of one or more quantities of interest. The optimal set of outputs yopt is defined as the most reliable output values and can either be (i) real data obtained from measurements or observations or (ii) synthetic data, that is, predicted values generated by a model that is considered sufficiently accurate by the user. For each set of parameters θ and one output (or prediction point) ym(θ), we define the residual rm(θ) as the difference between the optimal value ymopt and the output ym(θ), that is,

rmθ=ymoptymθ.(1)

Table 1
www.frontiersin.org

Table 1. Notation used in the FISR method description.

In addition, we define the cost function as

Cθ=m=1Mymoptymθ2.(2)

The cost quantifies how far the outputs obtained with parameter set θ, i.e., y(θ), are from the optimal yopt and is therefore used to assess the reliability of a parameter set θ. Furthermore, the sensitivity of the model residuals to changes in the parameters is arranged in the Jacobian matrix J, with elements

Jmμ=rmθμ.(3)

Finally, the Fisher Information Matrix (FIM) g has a key role in our method, and its elements are defined as

gμν=m=1Mrmθμrmθν.(4)

In matrix form, the FIM and the Jacobian are related through g=JTJ, being J (Equation 3) a matrix of size M×N and g (Equation 4) a matrix of size N×N. As the FIM g is square, symmetrical and positive semi-definite, all its eigenvalues are real and non-negative, and quantify how strongly perturbations along the corresponding eigenvector directions in the parameter space affect the model output y. Note that, as the Jacobian represents the change in the output with respect to parameter perturbations, it is a matrix of sensitivities (Saltelli, 2008). Thus, another interpretation of the FIM is the matrix of second-order sensitivities (also known as parameter synergies) (Faro et al., 2019).

In cases where the parameters are positive by definition and/or their scales are significantly different, it will be very helpful to calculate J and g after a re-parametrization of the model such that θ̃μ=ln(θμ), as was already stated in the original MBAM algorithm (see, for example, Transtrum and Qiu (2016)). This re-parametrization is aimed at weighing the order of magnitude of the parameter rather than its value, so the method does not necessarily bias the elimination procedure toward smaller values.

In practice, sloppy models are those whose FIM eigenvalues span along many orders of magnitude, from eigenvalues close to zero, whose associated eigenvectors pinpoint the parameter directions that hardly affect the system—the sloppy directions—, to large eigenvalues, whose eigenvectors denote directions of high parameter sensitivity. We will exploit this feature of the FIM for parameter reduction in the next Subsection.

2.2 Parameter reduction algorithm

The FISR algorithm is implemented through successive dimensionality reduction steps, reducing the original model iteratively until no further simplifications are possible. Each reduction step takes an initial model as the input and produces a simplified version as the output. This iterative process is carried out by an algorithm structured in the following three stages:

Stage 1. Navigation through the parameter space

Considering an initial model y(θ), each set of parameters θ represents one point in the parameter space associated with y(θ). This set has an associated FIM g(θ) and cost C(θ). The parameter navigation consists of a path parameterized by the variable τ, θ(τ), that starts from the initial point θ(τ0) (where the cost is minimum), and evolves according to

θτi+1=θτi+vτiδτ,(5)

where v(τi) is the velocity vector at iteration i and v(τi)δτ denotes a small size step. The velocity v(τi) is a L2-normalized vector in the direction of the eigenvector corresponding to the smallest eigenvalue of the FIM g(θ(τi)) (namely, the sloppiest direction). Since the orientation of the eigenvectors is undefined, among the two possible ones, in each step τi the orientation of v(τi) is chosen such that it verifies that the angle of the velocities along the geodesic between consecutive steps is lower than π/2, i.e., that v(τi)v(τi1)>0, ensuring that the path θ(τ) does not go backward locally and therefore allowing a fast displacement from the original point1. This definition of v(τi) based on the FIM certifies that the small step v(τi)δτ has the minimum possible effect on the model output and thus on the cost. In principle, the smoothness of the model and the lack of global conserved quantities guarantee that the path θ(τ) reaches a boundary (Transtrum and Qiu, 2014). For instance, if a parameter is a positive real number, the theoretical limits would be zero and infinity (Transtrum et al., 2011), so the manifold boundary would correspond to a region where one or more parameters take those limiting values. In practice, the path is considered finished when the tendency of at least one parameter toward a limit is persistent over time. For consistency, we denote θ(τf) as the set of parameters after the last iteration of Equation 5.

Stage 2. Limit evaluation and generation of a new model

As mentioned above, stage 1 ends when one or more θμ reach a boundary (limiting value). In stage 2, a new model is constructed by considering the physical meaning of the limiting values of these parameters. For the typical case where parameters are positive real numbers and only one or two parameters reach a boundary, the possible scenarios are:

(i) A parameter approaches zero in stage 1. This may indicate that the microscopic process it represents has become negligible in the system with the new values of the parameters, for example, because it is much slower than the rest of the processes. In such cases, the new model excludes this process and the corresponding parameter.

(ii) A parameter approaches infinity in stage 1. This is typical when the parameters are reaction rate coefficients, and implies that a mechanism is almost instantaneous in comparison to others. This provides an adiabatic elimination of the fast timescale, and this mechanism can now be described by a single process with the proper change in the initial conditions.

(iii) Two or more parameters tend to infinity or one to infinity while the other to zero in stage 1. Here, a new parameter that combines them but remains finite is introduced in the model.

Note, however, that more complicated scenarios are possible, and more than two parameters may tend to zero or infinity in the same reduction step. In any case, the interpretation of the limit to generate the new model is a task that cannot be automated and must be done by hand because it requires knowledge about the processes that are modeled and the physical meaning of the parameters. Note that the new model is now evaluated in the parameter set θ, which is obtained by implementing the changes in θ(τf) and therefore has fewer parameters than the initial model.

Stage 3. Fitting of the new model

Once the new model is defined, it is fitted by finding the parameter set θ that minimizes the cost. The initial guess is θ. The outcome of stage 3 is a reduced model with fewer parameters than the initial one, but due to the methodology presented in stages 1, 2, and 3, the new reduced model achieves predictions as close as possible to those of the initial model.

Consecutive dimensional reduction steps. The whole process of parameter reduction starts with the original model y0 and is iterated several times, where each loop is defined as a dimensional reduction step s. The complete algorithm structure is shown in Figure 1. At each step s, model ys1 is reduced, yielding model ys2. A reduction step can be performed as long as there is a sloppy direction. We know that no further reductions are possible when the cost becomes so large that the output of the simplified model is too different from that of the original model. The minimal model will be the one that retains its predictive power with the smallest possible number of parameters.

Figure 1
Flowchart depicting a three-stage iterative algorithm for model reduction. Stage 1 involves navigating through parameter space, represented by (θ(τ)). Stage 2 leads to an unfitted model after interpretation of the limit. Stage 3 involves fitting to obtain a reduced model, resulting in (θ). The process iterates for subsequent reductions.

Figure 1. Structure of the FISR algorithm to reduce model complexity. The algorithmic flow can be divided in three stages: (stage 1) navigation through the parameter space of the initial model, constructing a path θ(τ) from the initial point θ(τ0) to the boundary θ(τf) using the eigenvector associated with the smallest eigenvalue of the Fisher Information Matrix g to guide the direction of change v(τi); (stage 2) limit evaluation that leads to the generation of a new model, where limits in θ(τf) are interpreted analytically to eliminate parameters and obtain θ; and (stage 3) fitting of the new model, minimizing the cost for the reduced model starting from θ to reach θ. The final reduced model of this process will serve as the initial model for the next dimensional reduction step, repeating the process until no further reductions are feasible, yielding the minimal predictive model.

3 Results

In this Section, we apply the FISR method introduced in this work to a phosphorus astrochemistry model comprising 14 chemical reactions involving 7 P-bearing molecules proposed by Fernández-Ruz et al. (2023). The parameters subject to reduction are the kinetic parameters, specifically the rate coefficients kμ of the 14 reactions. Our focus is on the abundances of PO and PN, the most abundant P-bearing molecules detected so far in star-forming regions (Lefloch et al., 2016; Rivilla et al., 2020; Fontani, 2024). We perform 9 dimensional reduction steps, 8 of which preserve with high precision the model’s ability to predict the evolution of PO and PN abundances within the time range from 104 to 105 yrs. These timescales are typical of molecular outflows in star-forming regions where PO and PN have been detected (e.g., Lefloch et al. (2016), Rivilla et al. (2020), Bernal et al. (2021), Fontani (2024)), and are the ones usually explored with astrochemical models (see e.g., Jiménez-Serra et al. (2018)). The FISR method results in a final minimal model with only 3 rate coefficients.

3.1 Original model for the chemical evolution of phosphorus in the ISM

We adopt the phosphorus chemistry model for the ISM from our recent publication (Fernández-Ruz et al., 2023) as the original model of the complexity reduction process. We opt for this model because it has been proven to accurately describe the [PO]/[PN] ratios of the whole phosphorus network predicted by UCLCHEM, while remaining sufficiently small to apply the FISR method. The system, whose associated chemical network is shown in Figure 2, describes the evolution of phosphorus chemistry in a star-forming region over 105 yrs. For the simulations, we assume constant physical conditions with a temperature of T = 100 K and a cloud H density of nH=104 cm3, typical of regions affected by shocks associated with molecular outflows and where PO and PN have been detected (see e.g., the works of Lefloch et al. (2016) or Rivilla et al. (2020)). The model represents the dynamics resulting from the application of the law of mass action to a network of 14 chemical reactions with their corresponding 14 reaction rate coefficients kμ, listed in Table 2, and involving 17 chemical species with initial conditions provided in Table 3. Among these species, there are 7 P-bearing molecules: P, PH, PH2, PH3, CP, PO, and PN. Importantly, most of the reaction rate coefficients kμ were calculated using the Arrhenius parameters extracted from the KIDA database (Wakelam et al., 2012). However, three of these coefficients were later adjusted for T = 100 K using Bayesian inference to obtain PO and PN predictions that match the observed values (see reactions 1, 2 and 10 in Table 2 and Fernández-Ruz et al., 2023). More precisely, the model predicts [PO]/[PN] = 3.3 at t=104 yrs, in agreement with Lefloch et al. (2016), Rivilla et al. (2018), Bernal et al. (2021), which reported [PO]/[PN] ratios equal to 2.8, 1.4 and 2.6, respectively. Finally, we assume that P is initially adsorbed onto dust grains, from which 1% is released into the gas phase (Jiménez-Serra et al., 2018; Fernández-Ruz et al., 2023). However, note that the actual depletion factor of P in the ISM is still a debated question (Fontani, 2024). In addition, we set a P-hydrogenation fraction fP of 0.5, meaning that 50% of the released P is hydrogenated, and the rest remains as atomic P. For simplicity, PH, PH2 and PH3 are formed in equal amounts, leading to the initial abundances shown in Table 3, but let us remark that there are currently no observational information about the abundance of P, PH, PH2 and PH3 in star-forming regions, and therefore, we can only provide reasonable guesses for these parameters. Interestingly, though, we have recently shown that the evolution of the system is in practice independent of the P-hydrogenation fraction fP for T = 100 K beyond 104 yrs, that is, in the time range of interest for this paper (Fernández-Ruz et al., 2023).

Figure 2
Diagram showing a complex network with circle nodes representing chemical species and square nodes (labeled R1–R14) denoting chemical reactions, with directed links pointing from reactants to reaction nodes, and from reaction nodes to products.

Figure 2. Bipartite chemical network representing the 14 chemical reactions involved in the phosphorus chemistry in the interstellar medium studied in this work. Orange circular nodes correspond to P-bearing species, whose abundance evolves over time as described by Equation 6. The rest of the species are represented as blue circular nodes. Square nodes (labeled R1–R14) denote chemical reactions, with directed links pointing from reactants to reaction nodes, and from reaction nodes to products.

Table 2
www.frontiersin.org

Table 2. Reactions and their rate coefficients. (a) The rate coefficients of these reactions were adjusted with Bayesian inference in Fernández-Ruz et al. (2023). (b) The Arrhenius parameters of these rate coefficients come from theoretical quantum chemical calculations.

Table 3
www.frontiersin.org

Table 3. Initial gas-phase abundances that apply in the original model, extracted from Fernández-Ruz et al. (2023). Some initial abundances are missing because they are not needed as they do not appear in the models’ equations. (a) In cases where the source provided two values or we considered two sources, we used the geometric mean. (b) Up to date, CP has not been detected in the ISM, but it has been detected in a circumstellar shell envelope by Guelin et al. (1990). Thus, in our model we consider that CP is present but we fix its initial value to 1013 so it is sufficiently below the detection limit (1012). (c) The value is manually set to one (for H) and zero (for PO and PN) in agreement with the models’ configuration. (d) As proposed in other works (Jiménez-Serra et al., 2018; Fernández-Ruz et al., 2023), we assume that P (cosmic abundance of 2.57×107) is depleted by a factor of 100 and that 50% is hydrogenated on the surface of dust grains forming PH, PH2 and PH3. The rest remains as atomic P. For simplicity, PH, PH2 and PH3 are formed in equal amounts, leading to the initial abundances shown in the table.

We refer to the system of ordinary differential equations (ODEs) with 17 variables, derived from applying the law of mass action to the astrochemical model plotted in Figure 2, as the total system. This system is mathematically intractable and can only be solved using numerical methods. However, the linear effective system is a linearized approximation of the full system that can be theoretically solved under certain approximations for the 7 P-bearing species. The linearization is possible when the non P-bearing species are treated as constants, an assumption justified by their significantly higher abundances, as shown in Table 3 (see Supplementary Section S1.1 for details). The linear effective system solution consists of explicit equations for the time-evolving abundances:

Xp=q=1p1Cpqrprqerqt+Cpperpt,(6)

where [Xp], with p=1 to 7, represents the abundance of the P-bearing species PH3, PH2, PH, CP, P, PO, and PN, respectively. The constants C and r depend on the reaction rate coefficients and initial abundances, with their explicit expressions provided in Supplementary Section S1.2. Also, note that, for p=1, the summation corresponds to an empty sum; thus, its value is zero. This theoretical solution of the linear effective system approximates the numerical solution of the total system for the P-bearing species with average relative errors of 1% for T = 100 K, while being the former more than 5 orders of magnitude computationally faster than the latter (Fernández-Ruz et al., 2023).

The sensitivity analysis presented in Fernández-Ruz et al. (2023) revealed that the abundances of PO and PN at times t=103, t=104, and t=105 yrs and for T = 100 K exhibit negligible dependence on some of the rate coefficients kμ and large sensitivity on a few of them. This indicates that the model shows strong parameter hierarchy or sloppiness, making it particularly well-suited for parameter reduction using the FISR method, with the rate coefficients kμ as the 14 parameters and PO and PN as the outputs. We do not consider the remaining P-bearing molecules as outputs because they have not been detected in star-forming regions (Fontani, 2024). Based on this, and following the notation defined in Section 2.1, we call original model—that will act as a model 0— to y=y(θ=k), where ym=ym(k1,,k14) and m=1,,50 enumerates the abundances of PO and PN given by Equation 6, evaluated at 25 time points uniformly distributed between t=104, and t=105 yrs—the time range typically observed in real data (see above). Finally, we need to obtain the optimal prediction yopt used in the calculation of the residuals and the cost (Equations 1, 2). Since we lack a temporal series of observational data points for the abundances of PO and PN in a molecular cloud—real data correspond to a single snapshot of the cloud’s time evolution—, we generate synthetic data from the original model, considering that yopt=y(koriginal), where koriginal is the set of rate coefficients values showed in Table 2. By construction, the cost of the original model, with koriginal as input, is zero.

3.2 Toward the simplest model compatible with the chemical evolution of phosphorus in the ISM

3.2.1 Dimensionality reduction steps

Here, we apply the FISR algorithm to the original model of the phosphorus chemistry in the ISM through a series of iterative dimensionality reduction steps. As explained in Section 2.2, each reduction step s in the algorithm consists of (i) the parameter space navigation in the model s1 until a boundary is reached, (ii) the evaluation of the limit and model reinterpretation to obtain model s, with less parameters, and (iii) the fitting of model s. Therefore, a new model with fewer parameters and, in some cases, fewer variables is derived at each reduction step s. We adopt the re-parametrization k̃μ=ln(kμ) justified in Section 2.1, and also the abundances of PO and PN are log-transformed in the calculation of the cost and the FIM. The fitting of models is done by finding the set of rate coefficients k that minimizes the cost with the Nelder-Mead method (Lagarias et al., 1998). In summary, our goal now is to evaluate the performance of the new models and find the simplest set of reactions that can yield time-evolution abundance curves of PO and PN from 104 to 105 yrs that are indistinguishable from those obtained from the original model. We remind the reader that we focus on timescales between 104 to 105 yrs because these are the relevant timescales for the star-forming regions where PO and PN have been detected (Lefloch et al., 2016; Rivilla et al., 2020; Fontani, 2024). Figure 3 shows the evolution of the set of rate coefficients k along the parameter space at each reduction step, obtained in stage 1 of the FISR algorithm. Each path has been computed for 105 iterations, far enough to appreciate a clear tendency in which at least one rate coefficient kμ goes to zero or infinity. The cases where a parameter kμ goes to zero are easy to interpret: the corresponding reaction can be removed from the model since the FISR algorithm proves that it has a negligible effect on the cost, which ultimately means that the reaction is not relevant for the final abundances of PO and PN. Therefore, the new model s has one less rate coefficient and is obtained by removing that reaction from model s1. On the contrary, the cases where a kμ goes to infinity must be interpreted individually.

Figure 3
Nine graphs illustrating the evolution with τ of ln(k). Graph (A) shows k5 approaching zero, (B) k7 to zero, (C) k11 to zero, (D) k8 to infinity, (E) k12 to infinity, (F) k13 to infinity, (G) k9 to zero, (H) k4, k10, k14 to infinity, and (I) k1 to zero. Each step details the corresponding change visually.

Figure 3. Evolution with τ of the rate coefficients kμ of the phosphorus astrochemical system in the ISM along the path θ(τ)=k(τ) obtained in stage 1 of the FISR algorithm, for 9 dimensionality reduction steps. τ is the time step in the navigation through the parameter space. The path starts at the initial point k(0) where the cost is minimum and tends to a boundary where one or more kμ reach zero or infinity. In steps 1, 2, 3, 7 and 9 (panels A, B, C, G, I), the rate coefficients k5, k7, k11, k9 and k1, respectively, approach zero. In steps 4, 5 and 6 (panels D, E, F), the rate coefficients k8, k12 and k13 tend to infinity. In step 8 (panel H), the rate coefficients k4, k10 and k14 tend to infinity in parallel. The number of iterations is 5×104 for all reduction steps except for step 5, which only reaches 7,000 iterations due to computational limitations arising from exceedingly large values.

After each reduction step, a new model is built, and this process is repeated along 9 dimensionality reduction steps. Figure 4 shows the time evolution of the abundances for the P-bearing species of every new model s compared to the original model. For completeness, Supplementary Section S2 shows the same comparison for the ratio [PO]/[PN], a relevant quantity in phosphorus astrochemistry, and Supplementary Section S3 compiles the initial conditions and chemical reactions that constitute every model, from the 14 reactions of the original model to the 2 reactions of model 9, the most simplified one.

Figure 4
Nine graphs labeled A to I display models 1 to 9, showing the abundances of chemical species over time in years. The x-axis represents time, ranging from 1 to 10^5 years, and the y-axis represents abundances from 10^(-13) to 10^(-8). Each graph includes multiple colored lines, with solid and dashed lines indicating different species and conditions, as described in the legend.

Figure 4. Evolution with time of the abundances relative to H of the P-bearing molecules for models 1 to 9 (solid lines), obtained in stage 3 of the FISR algorithm, and for the original model (O.M., dashed lines; Fernández-Ruz et al., 2023) of the phosphorus astrochemistry in the ISM. Models 1–9 (panels A–I) are progressively simpler approximations of the original model. Although only abundances of PO and PN beyond 104 yrs (vertical black lines) were used as input in the reduction algorithm, models 1-8 reproduce precisely the abundances of PO and PN of the original model beyond a few 100 years. On the contrary, the disagreement between model 9 and the original model can be seen with the naked eye.

As we can see in Figures 3A–C, corresponding to reduction steps 1, 2 and 3, respectively, the reaction rate coefficients k5, k7 and k11 tend to zero. In consequence, reactions 5 (P+O2 PO + O), 7 (N + PH PN + H) and 11 (O + CP P + CO) can be interpreted as infinitely slow in comparison to the rest, and can be sequentially removed, yielding models 1, 2 and 3. For reaction 5, this is consistent with new quantum chemical calculations of this rate constant, which show that its value at 100 K is very small (de la Concepción et al., 2024). In Figures 4A–C we see that PN and CP curves differ between the original model (O.M., dashed lines) and the new models (solid lines) for low times, but the PO and PN time-evolution abundance profiles after t=103 yrs remain indistinguishable. Noticeably, reduction step 1 shows a strongly noisy behavior for k5 because the equations of the original model present a structural non-identifiability regarding parameters k5 and k6, as the mathematical solutions of the system depend on these two parameters exclusively through the sum k5[O2]0+k6[OH]0. Structural non-identifiability arises when the model’s equations contain redundant parameters or symmetries, making it impossible to uniquely determine two or more parameters from the data, regardless of its quality.

Figure 3D shows that the rate coefficient of reaction 8 (N + CP PN + C) approaches infinity in the reduction step 4. This means that we can define a new model—model 4, see Supplementary Section S3 —where this reaction is infinitely fast in comparison to the rest, so the initial abundance of CP is added to the initial abundance of PN, removing CP as a variable of the system3. Besides reaction 8, only reaction 14 (C + PH CP + H) involves CP in model 3, so it can be merged with reaction 8, yielding the new reaction 14 (C + PH + N PN + H + C) that accounts for the 2-step process: C + PH CP + H followed by N + CP PN + C. Since in this case the first reaction is the rate-determining step, the rate law of the new reaction 14 is v=k14[C][PH]. Figure 4D shows that, despite model 4 lacking CP (green solid line is absent), new reaction 14 can capture the growth in the abundance of PN from the original model.

In reduction step 5, the rate coefficient of reaction 12 (H+PH2 PH+H2) tends to infinity, as shown in Figure 3E. This indicates that PH2 is consumed much faster (instantaneously relative to the time granularity of the data) than the rest of the species. Therefore, model 5 can do without PH2 as a variable and the initial abundance of PH2 can be added to the initial abundance of PH. Reaction 13 (H+PH3 PH2+H2) produces PH2, so it can be merged with reaction 12 to yield 2H+PH3 PH+2H2, which becomes new reaction 13. Similarly to the reduction step 4, the rate law is derived from the rate-determining step, thus for the new reaction 13, the rate law is v=k13[H][PH3]. In addition, since we are assuming that PH2 is consumed immediately in reaction 12, reaction 3 (O+PH2 PO+H2) is also erased. Figure 4E shows that after these changes are applied, the evolution curves of model 5 remain practically equivalent to those of the original model but with the absence of CP and PH2.

Similarly to the former two steps, in the reduction step 6 (Figure 3F), the rate coefficient k13 of reaction 2H+PH3 PH+2H2 approaches infinity. In model 6, the initial abundance of PH3 can, therefore, be added to the initial abundance of PH, removing variable PH3. This reaction is not merged with any other reaction because no reaction forms PH3, thus it is simply removed in model 6. Figure 4F shows that in this new model, PO and PN do not match the original curves for short times, but they do so for times longer than some 100 years. Interestingly, although PH3 disappears, the evolution of PH does not resemble that of the original PH and, on the contrary, plays a similar role to the original PH3. At this stage of the dimensional reduction progression, the chemistry exerted by PH, PH2 and PH3 has been condensed in the evolution of a single species, PH.

The reduction step 7 is described in Figure 3G, where reaction rate coefficient k9 tends to zero, and therefore reaction 9 (P + CN PN + C) can be eliminated. Figure 4G shows that this leads to a time shift with respect to model 6 in Figure 4F: PH abundance decays over longer timescales in model 7 than in model 6, while PO and PN reach detectable abundances (1012 relative to H) later in model 7 than in model 6. This represents a significant change in the new model for t<103 yrs, although between t103105 yrs the abundance evolution profiles of P, PO and PN remain indistinguishable from those of the original model. In summary, reaction 9 plays an important role in the dynamics at short times, but its omission does not affect the dynamics of PO and PN within the time frame of interest.

Interestingly, the reduction step 8 shown in Figure 3H involves 3 parameters. The rate coefficients of reactions 4 (O + PH PO + H), 10 (H + PH P+H2), and 14 (C + N + PH PN + C + H) approach infinity at the same speed. These reactions consume PH and transform it into PO, P, and PN, respectively. In the new model, these reactions are instantaneous, and the initial abundance of PH can be distributed into the initial abundances of PO, P, and PN, as shown in Supplementary Section S4. These changes yield model 8 that, as can be seen in Figure 4H, only has P, PO and PN as variables. Notably, at this stage of complexity reduction, the system still predicts PO and PN abundances beyond t=103 yrs that are identical with the naked eye to those of the original model.

The last iteration of the FISR algorithm, reduction step 9, is shown in Figure 3I. Reaction rate coefficient k1 tends to zero, and consequently reaction 1 (N + PO P + NO) is eliminated. Only reactions 2 (N + PO PN + O) and 6 (P + OH PO + H) remain. This is the first model with visible differences between its PO and PN curves and those of the original model for t>104 yrs (Figure 4I). Anyway, while model 9 still predicts final abundances of PO and PN close to the original ones, a new reduction step—as described in Supplementary Section S5 —, yields a system with a unique reaction and PO and PN curves completely different from the original model for all times, clearly warning that the model cannot be simplified beyond model 9.

3.2.2 Predictive power and parameter sloppiness of the reduced models

As shown in the previous Subsection, at least one parameter tends toward an extreme value during each successive navigation through the multi-dimensional parameter space. However, it is important to note that the remaining parameters also undergo minor changes. This occurs because the vector v(τi) can point in any direction during each path iteration through the parameter space. Additionally, the values of all parameters may change when the new model is fitted. Figure 5A quantifies this phenomenon by showing the evolution of the set of rate coefficients, k, throughout all dimensional reduction steps, both after each parameter navigation and after each parameter fitting. Notably, the rate coefficients that persist at step 8 (k1, k2, and k6) exhibit minimal variation −0.3%— compared to their original values from the KIDA database and the work by García de la Concepción et al. (2021).

Figure 5
Chart (A) shows logarithmic values of fourteen parameters (\(k_1\) to \(k_{14}\)) over nine reduction steps, with periodic parameter navigation (P.N.) and fitting (Fit). Chart (B) illustrates cost fluctuations of a process across the same steps, with visible peaks and troughs. Chart (C) presents FIM eigenvalues for models with varying parameters, labeled from Model 0 (fourteen parameters) to Model 9 (two parameters), showing distribution of eigenvalues for each model.

Figure 5. Evolution of the parameters, cost, and Fisher Information Matrix spectrum throughout the 9 dimensional reduction steps of the FISR algorithm applied to the phosphorus astrochemical system in the ISM. (A) Values of the rate coefficients at the parameter navigation (P.N.) and fitting of the new model (Fit.) (stages 1 and 3 of the FISR algorithm, respectively) for each reduction step. (B) Cost of the corresponding models after the parameter navigation stage and the fitting stage of the FISR algorithm. A dashed horizontal line is displayed at cost C=103 to represent a limit for tolerable cost. (C) Fisher Information Matrix spectrum for each model s, i.e., the magnitude of all the eigenvalues of the FIM evaluated at the initial point of each navigation step.

Figure 5B shows the cost—calculated using Equation 2— after each parameter navigation and parameter fitting step. Since the cost associated with model s quantifies the difference between its predictions and those of the original model, low-cost values indicate that the reduction steps preserve the model’s predictive power. While the FISR algorithm identifies the path that minimizes the cost increase, some growth in cost during the parameter navigation stage is inevitable. However, the fitting process consistently offsets this increase to some extent. Consistent with the observation that the abundance plots of PO and PN at long timescales remain indistinguishable from the original curves for models 1–8, the cost during these steps remains very low, reaching 6×107 at the end of step 8. In contrast, the cost sharply rises to 0.55 by the end of step 9.

Finally, Figure 5C displays the eigenspectrum of each model, i.e., the set of eigenvalues of the Fisher Information Matrix g evaluated at the initial point k(τ0) of each navigation step through the parameter space. Since each eigenvalue quantifies the model’s sensitivity along the direction of its corresponding eigenvector, and the eigenvalues of the original model span 22 orders of magnitude, this plot highlights the significant sloppiness and parameter hierarchy inherent in the current models used to study the chemical evolution of phosphorus in the ISM.

3.3 Final model for the chemical evolution of phosphorus in the ISM

The FISR method applied in this study enabled the reduction of the original 14-reaction model to significantly simpler models that still accurately reproduce the chemical evolution of PO and PN in the ISM beyond t103 yrs—and note that only values beyond 104 yrs were used as input in the reduction algorithm. The algorithm was initiated with rate coefficients calculated for a temperature of T = 100 K, but note that the results presented here are not restricted to a single temperature but are more general (see the Discussion and Supplementary Section S6 for details). Notably, after eight reduction steps, the chemical kinetics was simplified to include only reactions 1 (N + PO P + NO), 2 (N + PO PN + O), and 6 (P + OH PO + H). Despite this significant reduction, the associated cost (C6×107) was so low that the PO and PN abundances beyond t103 yrs were indistinguishable from those predicted by the original 14-reaction model. As a result, model 8 provides a simplified and interpretable framework that accurately reproduces the long-term evolution of PO and PN formation predicted by the original model. As illustrated in Figure 4H, P acts as a reservoir due to its much higher abundance than other species. Since P is directly converted into PO through reaction 6, PO becomes more abundant than PN during the early stages of molecular cloud evolution. However, PO is irreversibly transformed into PN via reaction 2, which accounts for the resulting curves where [PN]>[PO] for t>3×104 yrs.

Model 8 was constructed through eight consecutive reduction steps. Interestingly, while its dynamics is derived from reactions 1, 2, and 6 with rate coefficients almost identical to those in the original model, the overall output is not solely explained by the activity of these reactions. It also incorporates the influence of reactions 4, 8, 10, 12, 13, and 14, all treated as infinitely fast, through modifications in the initial conditions of PH, P, PO, and PN in model 8. These adjustments account for the reactants transformed into products at much faster rates than the remaining relevant reactions. Essentially, these fast processes involved the cascade of PH3 PH2 PH, the conversion of PH into P, PO, and CP, and the transformation of CP into PN. The elimination of these reactions did not significantly increase the cost, as they occur on much shorter timescales than reactions 1, 2, and 6. Consequently, two distinct stages can be identified in the phosphorus chemical evolution under study: an early stage (up to t103 yrs) and a late stage (from t103 to 105 yrs). Model 8 effectively describes the dynamics of P, PO, and PN during the late stage, which accounts for the majority of the simulation time, while accurately capturing the final outcomes of the early chemistry without the need to reproduce its detailed dynamics. A schematic representation of the processes governing the chemistry in both stages is shown in Figure 6.

Figure 6
Chemical reaction network diagram displaying transformations among compounds P, PO, PN, PH, PH₂, PH₃, and CP. Arrows indicate reaction directions with rate constants k₁ to k₁₄. The network is divided into early stage (until 1,000 years) and late stage (1,000 to 100,000 years).

Figure 6. Schematic representation of the evolution of phosphorus through the chemical processes leading to the formation of PO and PN in the ISM across two distinct time stages. During the early stage (up to 103 yrs), PH forms via the dissociation of PH2 and PH3, subsequently reacting with O, H, and C to produce P, PO, and CP, respectively. Finally, CP is converted into PN through its reaction with N. In the late stage (103105 yrs), P and PO interconvert through reactions with OH and N, respectively, while PN forms via the reaction of PO with N.

With one additional reduction step, reaction 1 is eliminated. Model 9 represents the direct transformation of P into PO and subsequently into PN, but with an associated cost of 0.55 that is 105 times the cost of the three-reaction model. This increased cost results in a noticeable discrepancy between the abundance evolution profiles of PO and PN in model 9 and those of the original model. When applying dimensionality reduction techniques, there is always a trade-off between a model’s simplicity and the accuracy of its results, with the choice of an appropriate minimal model depending on the specific context and requirements. In this case, model 8 provides highly accurate results (as evidenced by its low cost) while remaining interpretable, making it effective for explaining the formation of PO and PN. However, model 9 offers an even simpler chemical network and may be suitable in situations where a higher cost is acceptable. Finally, further parameter reduction—step 10— leads the system to a single reaction, N + PO PN + O, with a cost that increases to approximately 160. At this stage, the model can no longer accurately reproduce the abundances of PO and PN, and this last step must be avoided. More information on how to obtain the minimal model with the FISR method and full details of the reduction process for step 10 are provided in Supplementary Section S5.

4 Discussion

In this work, we introduce a novel dimensionality reduction technique called the Fisher Information Spectral Reduction (FISR) method. It measures the sensitivity of the output on the parameters in any multi-parameter model, and we used it to reduce the complexity of a system describing the chemical network of phosphorus in the interstellar medium. This P-chemistry model, previously studied in Fernández-Ruz et al. (2023), demonstrates a strong parameter hierarchy, with high dependence on a few parameter directions, while exhibiting a pronounced insensitivity to the rest—commonly referred to as stiff and sloppy directions, respectively. Using the FISR method, we have exploited this property showing that a kinetic model with 3 chemical reactions (N + PO P + NO, N + PO PN + O, and P + OH PO + H, corresponding to model 8 in Section 3) and their 3 associated rate coefficients can predict time-evolution abundance profiles of PO and PN in the time range between 103 and 105 yrs that are indistinguishable—with a cost or error of approximately 107— from those of the original model, which consisted of 14 chemical reactions and 14 rate coefficients. Notably, the rate coefficients associated with these 3 chemical reactions remain nearly unchanged throughout the reduction process, suggesting that even subtle changes in these rate coefficients could drastically affect the system’s predictions. In fact, their values differ by less than 0.3% between the original and any of the reduced models, while the uncertainties associated with them in the KIDA database correspond to a 1σ confidence interval bounded by k/2 and 2k (Wakelam et al., 2012). In consequence, as a key takeaway for observational and theoretical astrochemists, our findings emphasize the importance of distinguishing between sloppy and stiff parameters in astrochemical models to target the key chemical reactions (the stiff parameters) whose rate coefficients need to be determined with high accuracy either via laboratory experiments or quantum chemical computations. A good example is the rate coefficient of the reaction P + OH PO + H, found to be essential in our model and also in reproducing the observed [PO]/[PN] ratios (Jiménez-Serra et al., 2018), and which has recently been determined by quantum chemical calculations (García de la Concepción et al., 2021). Other works have recently explored the rate coefficients of the reactions N + PO P + NO and N + PO PN + O (Douglas et al., 2022), although the level of theory of these quantum chemical computations is low.

We began this article by outlining the various challenges that phosphorus astrochemistry and its connection to the origin of life on Earth still present. Let us now examine how the drastic simplification of the phosphorus astrochemical network to its fundamental core, as presented here, enables a deeper understanding of its chemistry. Despite the minimal model 8 having only 3 rate coefficients as inputs, it accounts for the effects of other reactions without explicitly including their rate coefficients in the equations. The reduction steps in which one or more rate coefficients tend to infinity are interpreted as the corresponding reactions being instantaneous, and this translates into appropriate changes in the initial conditions. As shown in Figure 4, in the original model (dashed lines), by t103 yrs all CP and PH are already depleted, in part due to the dehydrogenation of PH2 and PH3. Therefore, the velocities of these processes are irrelevant for the long-term dynamics of P, PO, and PN as long as the associated reactions have reached equilibrium by t=103 yrs. Based on this, we were able to explain PO and PN formation as a two-step process: (i) an early stage, that can be considered infinitely fast; and (ii) a late stage, whose dynamics must be explicitly modeled with reactions 1, 2 and 6. The 2-stage interpretation of P-chemistry in the ISM implies that the exact values of the rate coefficients for the reactions operating on short timescales are not necessary to accurately predict the abundance evolution profiles of PO and PN on long timescales. However, the counterpart is that nothing about the early chemistry (involving PH, PH2, PH3 and CP) can be inferred solely from the PO and PN detections, since we demonstrated that the observed abundances of PO and PN do not serve to constrain the values of the rate coefficients of the early stage. A more complex description of P-chemistry on short timescales in star-forming regions would require a more complete set of observational data, including, for example, the abundance of PH3 in the interstellar medium (especially toward the same regions where PO and PN have been detected). PH3, however, remains elusive (Agúndez et al., 2008). Recent high-sensitivity observations carried out with ALMA toward dark molecular clouds have not yielded any detection (upper limit to the PH3 abundance of 5×1012; Rivilla et al. (2018), Furuya and Shimonishi (2024)), which may be due to: i) PH3 not being sufficiently abundant in the gas phase (e.g., it may be heavily frozen-out onto ices or converted very rapidly into other P-bearing molecules), or ii) PH3 being very difficult to observe from the ground given that its lowest energy transitions appear at sub-millimeter wavelengths (Agúndez et al., 2008).

It is worth noting that, although the FISR algorithm was applied using rate coefficients calculated for T = 100 K, the resulting reduced models remain good approximations (i.e., with low associated cost) over a certain temperature range. This is due to the system’s inherent sloppiness, which confers robustness to variations in the rate coefficients and, consequently, to changes in temperature. In Supplementary Section S6, we analyze the generalization of models 1–9 —originally obtained at T = 100 K—to temperatures ranging from 40 K to 160 K, without reapplying the full algorithm at each temperature. This robustness is advantageous, as it removes the need to re-run the computationally expensive FISR algorithm for every temperature. We find that the dynamics of phosphorus chemistry in the 40–160 K range can still be interpreted as a two-stage process, being the approximation of the early stage as instantaneous particularly accurate at higher temperatures. At very low temperatures, applying the full FISR method becomes necessary to obtain precise predictions. Accordingly, Supplementary Section S8 shows the costs of models 1–9 at T = 40, 70, 100, 130, and 160 K.

Relative to other dimensionality reduction methods currently available in the literature, the FISR method proposed here is significantly simpler. Specifically, in comparison to the MBAM algorithm (Transtrum and Qiu, 2014), upon which it is based, the FISR algorithm employs a streamlined methodology that is easier to implement, conceptually simpler, and admits larger output vectors (see Supplementary Section S7). Nevertheless, it is important to note that neither the FISR method nor the MBAM is currently suitable for application to very large systems due to several limitations: (i) both are based on a streamlined algorithmic structure that cannot be parallelized, (ii) each point of the trajectory navigating the parameter space requires the diagonalization of the Fisher Information Matrix computed numerically from the solution of the ODEs—in the phosphorus network, we were able to solve the model analytically, but this will not generally be feasible—and (iii) manual analysis of each reduced model is required at the end of each reduction step. In summary, it is crucial to continue advancing this promising line of research by developing new algorithms capable of analyzing more complex systems. Such advancements will enable a comprehensive study of the chemistry of the interstellar medium, where we believe parameter sloppiness is the rule rather than the exception.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.

Author contributions

MF-R: Conceptualization, Formal Analysis, Investigation, Methodology, Software, Writing – original draft, Writing – review and editing, Data curation, Visualization. IJ-S: Conceptualization, Investigation, Methodology, Writing – original draft, Writing – review and editing. MC: Conceptualization, Investigation, Methodology, Writing – original draft, Writing – review and editing, Formal Analysis, Software. MR-B: Investigation, Writing – original draft, Writing – review and editing, Visualization. JA: Investigation, Writing – original draft, Writing – review and editing, Conceptualization, Formal Analysis, Funding acquisition, Methodology, Project administration, Software, Supervision.

Funding

The author(s) declare that financial support was received for the research and/or publication of this article. JA and MF-R received support from grant No. PID2021-122936NB-I00, MC from grant No. PID2022-140217NB-I00, MR-B from grant PID2022-140180OB-C22, and IJ-S from grant PID2022-136814NB-I00, all of them funded by the Spanish Ministry of Science and Innovation/State Agency of Research MCIN/AEI/10.13039/501100011033 and by “ERDF A way of making Europe''. JA acknowledges financial support from grant No. PIE2024ICT085 and IJ-S from grant No. PIE 202250E155, both of them funded by the Spanish National Research Council (CSIC). MR-B is supported by the Instituto Nacional de Técnica Aeroespacial ”Esteban Terradas'' (INTA). MF-R is supported by a predoctoral contract from INTA. Authors benefited from the interdisciplinary framework provided by CSIC through the “LifeHUB.CSIC” initiative (PIE 202120E047-Conexiones-Life). IJ-S also acknowledges funding from the ERC grant OPENS (project number 101125858) funded by the European Union. Views and opinions expressed are however those of the authors only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible for them.

Acknowledgments

The authors are indebted to S. Ares, J. A. Cuesta and M. Martínez-Jiménez for fruitful comments.

Conflict of interest

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

The reviewer FF declared a past co-authorship with the author(s) IJ-S to the handling editor.

Generative AI statement

The authors declare that Generative AI was used in the creation of this manuscript exclusively to review the English precision of some sentences in the manuscript.

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.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fspas.2025.1570525/full#supplementary-material

Footnotes

1Note that with this criterion, the orientation of v(τ0) remains undefined, meaning that two valid paths coexist and one can choose any of them arbitrarily.

2From now on, we will refer to model ys as model s for simplicity.

3Recall that the non P-bearing species are assumed to be constant due to their higher abundances and the P-bearing species become the only dynamical variables. Therefore, for all the cases in this Subsection, we only consider changes in the initial conditions of the P-bearing species, as the changes in the non P-bearing species are negligible.

References

Agúndez, M., Cernicharo, J., and Guélin, M. (2007). Discovery of phosphaethyne (HCP) in space: phosphorus chemistry in circumstellar envelopes. Astrophysical J. Lett. 662, L91–L94. doi:10.1086/519561

CrossRef Full Text | Google Scholar

Agúndez, M., Cernicharo, J., Pardo, J. R., Guélin, M., and Phillips, T. G. (2008). Tentative detection of phosphine in IRC +10216. Astronomy Astrophysics 485, L33–L36. doi:10.1051/0004-6361:200810193

CrossRef Full Text | Google Scholar

Agúndez, M., and Wakelam, V. (2013). Chemistry of dark clouds: databases, networks, and models. ChRv 113, 8710–8737. doi:10.1021/cr4001176

CrossRef Full Text | Google Scholar

Antoulas, A. C. (2005). Approximation of large-scale dynamical systems. Philadelphia: SIAM.

Google Scholar

Aota, T., and Aikawa, Y. (2012). Phosphorus chemistry in the shocked region L1157 B1. Astrophysical J. 761, 74. doi:10.1088/0004-637X/761/1/74

CrossRef Full Text | Google Scholar

Bernal, J. J., Koelemay, L. A., and Ziurys, L. M. (2021). Detection of PO in orion-KL: phosphorus chemistry in the plateau outflow. Astrophysical J. 906, 55. doi:10.3847/1538-4357/abc87b

CrossRef Full Text | Google Scholar

Brown, K. S., and Sethna, J. P. (2003). Statistical mechanical approaches to models with many poorly known parameters. Phys. Rev. E 68, 021904. doi:10.1103/physreve.68.021904

CrossRef Full Text | Google Scholar

Chantzos, J., Rivilla, V. M., Vasyunin, A., Redaelli, E., Bizzocchi, L., Fontani, F., et al. (2020). The first steps of interstellar phosphorus chemistry. Astronomy Astrophysics 633, A54. doi:10.1051/0004-6361/201936531

CrossRef Full Text | Google Scholar

Charnley, S. B., and Millar, T. J. (1994). The chemistry of phosphorus in hot molecular cores. MNRAS 270, 570–574. doi:10.1093/mnras/270.3.570

CrossRef Full Text | Google Scholar

Cooper, G. W., Onwo, W. M., and Cronin, J. R. (1992). Alkyl phosphonic acids and sulfonic acids in the murchison meteorite. Geochimica Cosmochimica Acta 56, 4109–4115. doi:10.1016/0016-7037(92)90023-C

CrossRef Full Text | Google Scholar

de la Concepción, J. G., Cavallotti, C., Barone, V., Puzzarini, C., and Jiménez-Serra, I. (2024). Relevance of the P+O2 reaction for PO formation in astrochemical environments: electronic structure calculations and kinetic simulations. Astrophysical J. 963, 142. doi:10.3847/1538-4357/ad1ffa

CrossRef Full Text | Google Scholar

Douglas, K. M., Gobrecht, D., and Plane, J. M. C. (2022). Experimental study of the removal of excited state phosphorus atoms by H2O and H2: implications for the formation of PO in stellar winds. Mon. Notices R. Astronomical Soc. 515, 99–109. doi:10.1093/mnras/stac1684

CrossRef Full Text | Google Scholar

Faro, J., Von Haeften, B., Gardner, R., and Faro, E. (2019). A sensitivity analysis comparison of three models for the dynamics of germinal centers. Front. Immunol. 10, 2038. doi:10.3389/fimmu.2019.02038

CrossRef Full Text | Google Scholar

Fernández-Ruz, M., Jiménez-Serra, I., and Aguirre, J. (2023). A theoretical approach to the complex chemical evolution of phosphorus in the interstellar medium. Astrophysical J. 956, 47. doi:10.3847/1538-4357/acf290

CrossRef Full Text | Google Scholar

Fontani, F. (2024). Observations of phosphorus-bearing molecules in the interstellar medium. Front. Astronomy Space Sci. 11, 1451127. doi:10.3389/fspas.2024.1451127

CrossRef Full Text | Google Scholar

Fontani, F., Mininni, C., Beltrán, M. T., Rivilla, V. M., Colzi, L., Jiménez-Serra, I., et al. (2024). The GUAPOS project: G31.41+0.31 Unbiased ALMA sPectral Observational Survey. IV. Phosphorus-bearing molecules and their relation to shock tracers. Astronomy Astrophysics 682, A74. doi:10.1051/0004-6361/202348219

CrossRef Full Text | Google Scholar

Fontani, F., Rivilla, V. M., Caselli, P., Vasyunin, A., and Palau, A. (2016). Phosphorus-bearing molecules in massive dense cores. Astrophysical J. Lett. 822, L30. doi:10.3847/2041-8205/822/2/L30

CrossRef Full Text | Google Scholar

Furuya, K., and Shimonishi, T. (2024). Deep search for phosphine in a prestellar core. Astrophysical J. Lett. 968, L19. doi:10.3847/2041-8213/ad50cc

CrossRef Full Text | Google Scholar

García de la Concepción, J., Puzzarini, C., Barone, V., Jimenez-Serra, I., and Roncero, O. (2021). Formation of phosphorus monoxide (PO) in the interstellar medium: insights from quantum-chemical and kinetic calculations. Astrophysical J. 922, 169. doi:10.3847/1538-4357/ac1e94

CrossRef Full Text | Google Scholar

Glover, K. (1984). All optimal Hankel-norm approximations of linear multivariable systems and their L-error bounds. Int. J. control 39, 1115–1193. doi:10.1080/00207178408933239

CrossRef Full Text | Google Scholar

Goldsmith, P. F., Liseau, R., Bell, T. A., Black, J. H., Chen, J.-H., Hollenbach, D., et al. (2011). Herschel measurements of molecular oxygen in Orion. Astrophysical J. 737, 96. doi:10.1088/0004-637x/737/2/96

CrossRef Full Text | Google Scholar

Gomes, A. C. R., Souza, A. C., Jasper, A. W., and Galvão, B. R. L. (2023). The P(4S) + NH(3Σ) and N(4S) + PH(3Σ)reactions as sources of interstellar phosphorus nitride. Astronomical Soc. Aust. 40, e011. doi:10.1017/pasa.2023.13

CrossRef Full Text | Google Scholar

Guélin, M., Cernicharo, J., Paubert, G., and Turner, B. E. (1990). Free CP in IRC +10216. Astronomy Astrophysics 230, L9–L11.

Google Scholar

Gugercin, S., and Antoulas, A. C. (2004). A survey of model reduction by balanced truncation and some new results. Int. J. Control 77, 748–766. doi:10.1080/00207170410001713448

CrossRef Full Text | Google Scholar

Halfen, D. T., Clouthier, D. J., and Ziurys, L. M. (2008). Detection of the CCP radical (X2 Π r) in IRC +10216: a new interstellar phosphorus-containing species. Astrophysical J. 677, L101–L104. doi:10.1086/588024

CrossRef Full Text | Google Scholar

Holdship, J., Viti, S., Jiménez-Serra, I., Makrymallis, A., and Priestley, F. (2017). UCLCHEM: a gas-grain chemical code for clouds, cores, and C-shocks. Astronomical J. 154, 38. doi:10.3847/1538-3881/aa773f

CrossRef Full Text | Google Scholar

Jiménez-Serra, I., Viti, S., Quénard, D., and Holdship, J. (2018). The chemistry of phosphorus-bearing molecules under energetic phenomena. Astrophysical J. 862, 128. doi:10.3847/1538-4357/aacdf2

CrossRef Full Text | Google Scholar

Lagarias, J., Reeds, J., Wright, M., and Wright, P. (1998). Convergence properties of the nelder--mead simplex method in low dimensions. SIAM J. Optim. 9, 112–147. doi:10.1137/S1052623496303470

CrossRef Full Text | Google Scholar

Larsson, B., Liseau, R., Pagani, L., Bergman, P., Bernath, P., Biver, N., et al. (2007). Molecular oxygen in the ρ Ophiuchi cloud. A&A 466, 999–1003. doi:10.1051/0004-6361:20065500

CrossRef Full Text | Google Scholar

Lefloch, B., Codella, C., Montargès, M., Vastel, C., Podio, L., Viti, S., et al. (2024). Spatial distributions of PN and PO in the shock region L1157-B1. Astronomy Astrophysics 687, A75. doi:10.1051/0004-6361/202245338

CrossRef Full Text | Google Scholar

Lefloch, B., Vastel, C., Viti, S., Jimenez-Serra, I., Codella, C., Podio, L., et al. (2016). Phosphorus-bearing molecules in solar-type star-forming regions: first PO detection. MNRAS 462, 3937–3944. doi:10.1093/mnras/stw1918

CrossRef Full Text | Google Scholar

Maciá-Barber, E. (2020). The chemical evolution of phosphorus: an interdisciplinary approach to astrobiology. Oakville, Ontario, Canada: Apple Academic Press.

CrossRef Full Text | Google Scholar

Moore, B. C. (1981). Principal component analysis in linear systems: controllability, observability, and model reduction. IEEE Trans. Automatic Control 26, 17–32. doi:10.1109/tac.1981.1102568

CrossRef Full Text | Google Scholar

Pilorget, C., Baklouti, D., Bibring, J.-P., Brunetto, R., Ito, M., Franchi, I., et al. (2024). Phosphorus-rich grains in Ryugu samples with major biochemical potential. Nat. Astron. 8, 1529–1535. doi:10.1038/s41550-024-02366-w

CrossRef Full Text | Google Scholar

Quinn, K. N., Abbott, M. C., Transtrum, M. K., Machta, B. B., and Sethna, J. P. (2023). Information geometry for multiparameter models: new perspectives on the origin of simplicity. Rep. Prog. Phys. 86, 035901. doi:10.1088/1361-6633/aca6f8

CrossRef Full Text | Google Scholar

Rivilla, V. M., Drozdovskaya, M. N., Altwegg, K., Caselli, P., Beltrán, M. T., Fontani, F., et al. (2020). ALMA and ROSINA detections of phosphorus-bearing molecules: the interstellar thread between star-forming regions and comets. MNRAS 492, 1180–1198. doi:10.1093/mnras/stz3336

CrossRef Full Text | Google Scholar

Rivilla, V. M., Fontani, F., Beltrán, M. T., Vasyunin, A., Caselli, P., Martín-Pintado, J., et al. (2016). The first detections of the key prebiotic molecule PO in star-forming regions. Astrophysical J. 826, 161. doi:10.3847/0004-637X/826/2/161

CrossRef Full Text | Google Scholar

Rivilla, V. M., García De La Concepción, J., Jiménez-Serra, I., Martín-Pintado, J., Colzi, L., Tercero, B., et al. (2022). Ionize hard: interstellar PO+ detection. Front. Astronomy Space Sci. 9, 829288. doi:10.3389/fspas.2022.829288

CrossRef Full Text | Google Scholar

Rivilla, V. M., Jiménez-Serra, I., Zeng, S., Martín, S., Martín-Pintado, J., Armijos-Abendaño, J., et al. (2018). Phosphorus-bearing molecules in the galactic center. MNRAS Lett. 475, L30–L34. doi:10.1093/mnrasl/slx208

CrossRef Full Text | Google Scholar

Ruaud, M., Wakelam, V., and Hersant, F. (2016). Gas and grain chemical composition in cold cores as predicted by the Nautilus three-phase model. Mon. Notices R. Astronomical Soc. 459, 3756–3767. doi:10.1093/mnras/stw887

CrossRef Full Text | Google Scholar

Rugel, M. R., Beuther, H., Bihr, S., Wang, Y., Ott, J., Brunthaler, A., et al. (2018). OH absorption in the first quadrant of the Milky Way as seen by THOR. Astronomy Astrophysics 618, A159. doi:10.1051/0004-6361/201731872

CrossRef Full Text | Google Scholar

Saltelli, A. (2008). Global sensitivity analysis: the primer. John Wiley and Sons.

Google Scholar

Scibelli, S., Megías, A., Jiménez-Serra, I., Shirley, Y., Bergner, J., Ferrer Asensio, J., et al. (2025). First detections of PN, PO, and PO+ toward a shocked low-mass starless core. Astrophysical J. Lett. 985, L25. doi:10.3847/2041-8213/add344

CrossRef Full Text | Google Scholar

Sil, M., Srivastav, S., Bhat, B., Mondal, S. K., Gorai, P., Ghosh, R., et al. (2021). Chemical complexity of phosphorous-bearing species in various regions of the interstellar medium. Astronomical J. 162, 119. doi:10.3847/1538-3881/ac09f9

CrossRef Full Text | Google Scholar

Tenenbaum, E. D., Woolf, N. J., and Ziurys, L. M. (2007). Identification of phosphorus monoxide (X2Πr) in VY Canis majoris: detection of the first P-O bond in space. Astrophysical J. 666, L29–L32. doi:10.1086/521361

CrossRef Full Text | Google Scholar

Tenenbaum, E. D., and Ziurys, L. M. (2008). A search for phosphine in circumstellar envelopes: PH3 in IRC +10216 and CRL 2688? Astrophysical J. 680, L121–L124. doi:10.1086/589973

CrossRef Full Text | Google Scholar

Transtrum, M. K., Machta, B. B., and Sethna, J. P. (2011). Geometry of nonlinear least squares with applications to sloppy models and optimization. Phys. Rev. E 83, 036701. doi:10.1103/PhysRevE.83.036701

CrossRef Full Text | Google Scholar

Transtrum, M. K., and Qiu, P. (2014). Model reduction by manifold boundaries. Phys. Rev. Lett. 113, 098701. doi:10.1103/PhysRevLett.113.098701

CrossRef Full Text | Google Scholar

Transtrum, M. K., and Qiu, P. (2016). Bridging mechanistic and phenomenological models of complex biological systems. PLOS Comput. Biol. 12, e1004915. doi:10.1371/journal.pcbi.1004915

CrossRef Full Text | Google Scholar

Turner, B. E., and Bally, J. (1987). Detection of interstellar PN: the first identified phosphorus compound in the interstellar medium. Astrophysical J. Lett. 321, L75. doi:10.1086/185009

CrossRef Full Text | Google Scholar

Wakelam, V., Herbst, E., Loison, J.-C., Smith, I. W. M., Chandrasekaran, V., Pavone, B., et al. (2012). A kinetic database for astrochemistry (KIDA). Astrophysical J. Suppl. Ser. 199, 21. doi:10.1088/0067-0049/199/1/21

CrossRef Full Text | Google Scholar

Walton, C., Ewens, S., Coates, J., Blake, R., Planavsky, N., Reinhard, C., et al. (2023). Phosphorus availability on the early earth and the impacts of life. Nat. Geosci. 16, 399–409. doi:10.1038/s41561-023-01167-6

CrossRef Full Text | Google Scholar

Ziurys, L. M. (1987). Detection of interstellar PN: the first phosphorus-bearing species observed in molecular clouds. Astrophysical J. Lett. 321, L81. doi:10.1086/185010

CrossRef Full Text | Google Scholar

Keywords: phosphorus astrochemistry, interstellar medium, astrobiology, dynamical systems, chemical reaction networks, kinetic parameters, complexity reduction, parameter sensitivity

Citation: Fernández-Ruz M, Jiménez-Serra I, Castro M, Ruiz-Bermejo M and Aguirre J (2025) Strong parameter hierarchy in the interstellar phosphorus chemical network. Front. Astron. Space Sci. 12:1570525. doi: 10.3389/fspas.2025.1570525

Received: 03 February 2025; Accepted: 10 July 2025;
Published: 30 July 2025.

Edited by:

David Benoit, University of Hull, United Kingdom

Reviewed by:

Kotomi Taniguchi, National Astronomical Observatory of Japan (NAOJ), Japan
Francesco Fontani, Osservatorio Astrofisico di Arcetri (INAF), Italy

Copyright © 2025 Fernández-Ruz, Jiménez-Serra, Castro, Ruiz-Bermejo and Aguirre. 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: Jacobo Aguirre, amFndWlycmVAY2FiLmludGEtY3NpYy5lcw==

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.