We propose non-parametric methods for both local and global sensitivity analysis of chemical reaction models with correlated parameter dependencies. The developed mathematical and statistical tools are applied to a benchmark Langmuir competitive adsorption model on a close packed platinum surface, whose parameters, estimated from quantum-scale computations, are correlated and are limited in size (small data). The proposed mathematical methodology employs gradient-based methods to compute sensitivity indices. We observe that ranking influential parameters depends critically on whether or not correlations between parameters are taken into account. The impact of uncertainty in the correlation and the necessity of the proposed non-parametric perspective are demonstrated.

Uncertainty quantification (UQ) is a critical tool to assess the predictive ability of a model. This information can be used to map the envelop of model predictions, to improve a model via error reduction methods, and to inform control and optimization strategies in system tasks.1–9 Sensitivity analysis (SA) is one of the most effective tools for identifying influential model parameters.8,10–12 The SA approaches are typically classified as local or global methods. Local sensitivity analysis (LSA) computes variability in model predictions due to infinitesimal perturbations in the model parameters.8 The resulting local sensitivity indices (LSIs) include gradient8 and information-based methods.13 LSIs have also been applied to system optimization and model calibration problems.14,15 Global sensitivity analysis (GSA) determines variability in model predictions over a range of parameters due to uncertainty in those parameters. GSA techniques include analytical, regression, screening and variance-based methods.8,16–27

While significant progress has been achieved in UQ and SA methods over the years, physical systems often exhibit correlated parameters. In recent work, we introduced such a mathematical framework and demonstrated the impact of correlations on a model predictive ability for a complex reaction network.22 Our ability to understand and improve methods relies on developing simple but physically sound models that we can analyze mathematically.

In this paper we propose non-parametric methods for the GSA of chemical reaction models with correlated parameter dependencies. A Langmuir bimolecular hydrogen/oxygen competitive adsorption model is employed as a benchmark to motivate and concretely illustrate the derivation and algorithmic aspects of the proposed method. This system describes the competitive adsorption of hydrogen and oxygen on a Pt(111) surface. Such systems are encountered in catalytic oxidation, such as emissions abatement, small scale power generation, fuel cells and batteries. Here, parameter correlations stem from correlated quantum-scale computational data calculated using Density Functional Theory (DFT). These correlations are assimilated into the model as an informed prior distribution for the model’s parameters. The use of non-parametric methods in modeling parameter uncertainty and understanding global sensitivity is necessitated by the limited availability of quantum-scale data. The proposed mathematical methodology employs gradient-based methods to compute correlative local/global sensitivity indices (LSI/GSI) to illustrate the relative effects of parameter perturbations, errors and uncertainties in model parameters. We show that the ranking of influential parameters depends critically on whether or not correlations between parameters are taken into account. The impact of uncertainty in the correlation on the LSI/GSI is also demonstrated. Finally, we show the necessity of the proposed non-parametric perspective by comparing with a parametric approach.

The paper is organized as follows: the general form of the studied predictive model and LSI/GSI definitions are briefly presented in the “Background on sensitivity analysis” section. A detailed description of the impact of parameter correlations on LSA/GSA is addressed in the “Parameter correlation” section. Then, we apply the proposed CLSA/CGSA in the case of correlated hydrogen/oxygen binding energies in the “Langmuir bimolecular-adsorption model” section. Finally, we discuss methods to implement the proposed correlative LSA/GSA methods, other than direct numerical integration or Monte Carlo, namely generalized Polynomial Chaos methods that account for non-parametric noise models and parameter correlations.

In this section an appropriate mathematical framework is discussed for sensitivity analysis. First, we consider an ensemble of models of the general form

(1)

where Π(x|λ) denotes the predictive forward mathematical models, i.e. the probability distribution function (PDF) of state X = x for fixed K dimensional model parameters λ=[λ1  λ2    λK]TΛ, and Λ presents the parameter space. The term p(λ) denotes the PDF of λ which contains knowledge of uncertainty in the model, i.e. once we have p(λ) we can generate ensembles of X’s for each λ. Note that X may represent a static random variable, a snapshot of the system at some fixed time, or an entire time-series for dynamics and λ may denote the model parameters or indexing of different models.

In our specific model, λ corresponds to the binding energy of atomic oxygen and hydrogen on a given metal surface. We look at the uncertainty in coverage (Π) given the binding energy and its associated uncertainties and correlations. Coverage can also depend on other quantities that could be represented by λ, such as binding site, inert species, surface defects, surface impurities, and surface temperature.28–30 The formalism is though general and beyond the binding energy and the isotherm. Other physical systems that follow the Π(x|λ)p(λ) relationship include the dependency of molecular frequency (Π) on coverage (λ)31 and forecasted temperature changes (Π), with CO2, methane, and other greenhouse gases (λ).32–34 

The system observable can be defined over all possible realizations of the state

(2)

where h(x) denotes a desired quantity. The correlations in the parameter vector λ are also included in p(λ) and are propagated into the state X through the predictive forward model of Π(x|λ). Finally, the averaged observable for the model can be defined by

(3)

Consider a general class of nonlinear models of the form

(4)

where f is an arbitrary scalar function. The (relative) LSI of f with respect to λ of (4) at the nominal value of λ* is

(5)

where

(6)

The LSI of (5) supplies useful sensitivity information in the case of almost certain parameters, i.e. for a relatively tight range of parameter values. To incorporate the knowledge of uncertain parameter distributions and provide sensitivity information over the entire range, we determine the relevant GSI by employing the partial derivative of the LSI as a basic building block to integrate the local sensitivities over the total range of parameter changes

(7)

where q denotes the type of required index (q = 1: improved Morris index, q = 2: asymptotic limit of the standard Morris index).

Note that the PDF, which incorporates knowledge of the λ distribution in the GSI of (7), must be identified subject to available experimental and/or simulation-based data. The possible correlations between the system parameters, which may be discovered during regression of the data in statistical models, must be encoded in p(λ). Such correlations play a deciding role in sensitivity analysis and their effects are quantified in the following sections.

Previously, LSA and GSA were treated for the case of independent parameters. To extend sensitivity analysis to models with correlated parameters, we partition the vector of parameters into two,

(8)

where λ=[λ1  λ2]T, Λ = Λ1 ⊕ Λ2, λ1 contains all independent parameters, and λ2 contains all dependent parameters. Parameters can be classified through their correlations, which are identified using experiments and/or computational tools for specific case studies, and by applying causality statistical methods. When λ1 and λ2 are correlated, perturbations in one parameter affect the other. Proper mathematical tools are needed to quantify parameter correlations and their impact on model reliability.

The correlation between λ1 and λ2 can be described by their joint probability distribution,

(9)

For the marginal distribution of λ1,

(10)

Identifying p1) and p21) in a systematic way is an essential step in our CGSA.

The joint probability distribution of p1, λ2) can either be built directly, with a sufficiently large ensemble of experimental and/or simulation-based data,22 or computed sequentially by marginalization according to (9). The latter approach requires building PDFs with data for both p21) and p1), followed by Monte Carlo sampling to calculate correlative indices. There are various linear regression (LR) methods that can identify the conditional PDF of p21); including deterministic (DLR), stochastic (SLR), and Bayesian (BLR).35,36 DLR yields a deterministic linear model whose LCSIs can be computed exactly, while its GSIs depend on the choice of p1). SLR uses a least squares model for p21), typically in a Gaussian form.36 Usually, there is not enough data for adequate fitting of a least squares model. BLR can bypass this shortcoming by putting a prior on the parameters in the linear fit.35 

Hierarchical or empirical Bayesian methods can identify the marginal PDF of p1) via deterministic linear or stochastic nonlinear fitting to the data. Bootstrapping does not require fitting but instead relies on simulation,36 which is particularly appropriate for problems with little data, by creating synthetic new samples using random re-sampling according to the actual distribution of the data.36 In this way we can enrich the histogram with more data points and obtain a Gaussian distribution.36 If histograms are too sparse, “smoothed bootstrapping” can be used. This method applies a kernel to data from a standard histogram. A Bayesian approach can be used to fit data to a well known distribution that “looks like” the histogram. “Looks like” means that we pick a family of well known distributions and match the first few moments with the corresponding moments of the data’s histogram, i.e. mean, variance, skewness, etc.36,37

We perform the correlative sensitivity analysis by focusing on λ1, while still accounting for the correlations with λ2

(11)

The correlative local sensitivity index (CLSI) at the nominal point λ1* is obtained similarly to (5) by direct differentiation,

(12)

The CGSI can then be formulated

(13)

by employing the CLSI of (12) as building blocks where q = 1 or q = 2.

For deterministic correlation where λ2 = g1), we can simplify the λ1-marginal PDF of (10) by considering p21) = δ(g1) − λ2),

(14)

where δ(⋅) denotes the standard Dirac function. Therefore, from (11), we have

(15)

and the CLSI can be simplified to the following form

(16)

The additional second term in the CLSI differs from uncorrelative LSI in that the derivative with respect to the parameter λ2 comes directly into play. The CGSI formulation for such a simplified case is

(17)

The implementation of the sampling algorithm used to compute the correlative local/global sensitivity index (CLSI/CGSI) is described in Section S5 of the supplementary material.

We consider a Langmuir bimolecular adsorption model which describes competitive dissociative adsorption of hydrogen (H2) and oxygen (O2) on a catalyst surface,

(18)

where H2 and O2 denote the hydrogen and oxygen molecules in the gas phase, 2* are two active sites on the metal surface, and H* and O* represent the adsorbed hydrogen and oxygen atoms on the surface, respectively. A schematic of this adsorption process is illustrated in Figure 1. The physical system is related to hydrogen oxidation in fuel cells and batteries.38–42 

FIG. 1.

Competitive adsorption of hydrogen and oxygen on a catalyst surface.

FIG. 1.

Competitive adsorption of hydrogen and oxygen on a catalyst surface.

Close modal

The coverages dynamics can be formulated by the following set of ordinary differential equations

(19)

where θH*0 and θO*0 represent the initial hydrogen and oxygen coverages, respectively. PH2 and PO2 are the partial pressures of the gas phase species,43 and we set PH2=1.01325×1010N/m2, PO2=1.01325×1050N/m2 in this paper.

The hydrogen and oxygen coverages at equilibrium are

(20)

where P is partial pressure and K=kadskdes is the equilibrium constant.44,45K is determined from DFT calculations.

By focusing on variations of binding energies and fixing the other parameters, the coverages are

(21)

where ΔEH and ΔEO are the binding energies of atomic hydrogen and oxygen to the surface. It is the effect that uncertainty and correlations in the binding energies have on the coverages that we explore below. The detailed formulas of θ^H* and θ^O* is derived in Section S1 of the supplementary material.

The Langmuir adsorption isotherm is strictly valid at low coverage with adsorption at a single site, which is our system of interest. For dissociative adsorption, which we study here, the Langmuir model requires adjacent empty sites on the catalyst surface. An ab-initio molecular dynamics (AIMD) study of hydrogen on Pd(100) showed that regardless of coverage, only two adjacent catalyst sites are necessary to dissociate hydrogen.46 Although applications of AIMD to heterogeneous catalysis are rapidly advancing, the computational cost is still prohibitive for it to be used in generating adsorption isotherms.47 Less computational intensive methods, such as Monte Carlo48 and molecular dynamics with force fields,49 are used instead to generate an isotherm.

Seller et al. have shown that, when combined with the Bragg-Williams coverage model, the Langmuir adsorption isotherm accurately recreates experimental isotherms for several systems.48 Furthermore, the same study found that the Langmuir adsorption isotherm with mean field treatment compares favorably with coverages predicted from lattice based grand canonical Monte Carlo (GCMC) simulations under certain conditions. A force field based molecular dynamics simulation of dimethyl methylphosphonate (DMMP) also supports the validity of the Langmuir model.49 

Electronic contributions to adsorption enthalpies are calculated with DFT, using the Vienna ab initio Simulation Package (VASP), version 5.3.2,50–53 with the plane wave basis set, PAW pseudopotentials,54,55 and periodic boundary conditions. Simulation parameters are similar to those used in our previous work.56 All VASP input files are created using the Atomistic Simulation Environment, an open-source python-based software program.57 We employ the PBE exchange-correlation functional with D3 dispersion corrections.58,59 Spin-polarized calculations are performed for molecules in a vacuum and systems containing Ni and Co. The first Brillouin zone is sampled using the Monkhurst-Pack (3x3x1) mesh.60 For the purposes of this work, the level of accuracy achieved using this mesh size was sufficient. Ionic force cut-off for all calculations is set to 0.05 eV/Å. In slab calculations, we use the (4x4) supercell containing four layers of atoms, with the positions of the bottom two layers fixed. We use an adsorbate coverage of 1/16 monolayers in all calculations.

The DFT dissociative adsorption energy for molecular oxygen on a metal surface is defined in Equation 22. A similar relation holds for dissociative adsorption of molecular hydrogen.

(22)

such that

(23)

In Equation 22, EO2 is the DFT energy of an O2 molecule in a vacuum, E* is the DFT energy of the pristine metal slab, EO* is the energy of the adsorbate-metal system, D0 is the gas-phase bond dissociation energy at 0 K, and EO is DFT adsorption energy of an oxygen atom. The calculated energies for a variety of metal surfaces and the resulting scaling relationship between electronic contributions of H and O dissociative adsorption energies are shown in Figure 2. We obtain a linear correlation between hydrogen and oxygen adsorption energies with a R2 value of 0.87, i.e.

(24)

where a = 2.51, b = −2.46 (eV).

FIG. 2.

Correlation between oxygen and hydrogen adsorption energies on close-packed metal surfaces as defined in (24). Adsorbed atomic species are assumed to occupy fcc hollow sites.

FIG. 2.

Correlation between oxygen and hydrogen adsorption energies on close-packed metal surfaces as defined in (24). Adsorbed atomic species are assumed to occupy fcc hollow sites.

Close modal

Vibrational contributions and other temperature effects to adsorption enthalpy and entropy are accounted for in calculations of adsorption equilibrium constants (see Section S1 of the supplementary material for details). Zero point energy (ZPE) corrections for gas phase H2 and O2 are calculated using their experimental vibrational frequencies.61,62 Kinetic energy loss upon adsorption is accounted for by using the ideal gas value of 32RT. Harmonic and rigid rotor approximations were utilized to account for vibrational and rotational degrees of freedom, respectively.63 Hessian matrices are constructed using 0.015 Å displacements in x, y, and z directions from adsorbate equilibrium positions.

In order to validate our computational setup and provide an error estimate, we compare the calculated and experimental adsorption enthalpies of oxygen and hydrogen on platinum in Table I.64–67 The DFT calculations reproduce the experimental data well.

TABLE I.

Experimental and DFT-calculated enthalpies of adsorption for atomic oxygen and hydrogen on Pt(111).

AdsorbateExperimental enthalpyDFT computed enthalpya
3.71 ±0.07beV 3.68 eV 
2.63c eV 2.69 eV 
AdsorbateExperimental enthalpyDFT computed enthalpya
3.71 ±0.07beV 3.68 eV 
2.63c eV 2.69 eV 
a

Atomic O and H energies are calculated from ZPE-corrected O2 and H2 energies in a vacuum and their dissociation energies at 0 K provided by NIST.66 The improved agreement of the Pt-O bond enthalpy is from the use of molecular oxygen as a reference and is a result of error-cancellation as the PBE functional overestimates the bond enthalpy for both O2 and Pt-O when compared to experimental values. The PBE functional has been found to over-estimate adsorption energies on metal surfaces by 0.6 eV on average.67 While improved agreement of energies with experiment can therefore be achieved by using O2 as a reference, trends across metals for adsorbates should be the same regardless of whether atomic O or molecular O2 is used as a reference.

b

Experimental enthalpy for atomic oxygen adsorption, as measured by calorimetry, corrected and extrapolated to zero coverage by Karp et al.64 

c

Experimental enthalpy for atomic hydrogen adsorption, determined by correcting its experimental heat of formation at the 1/16 monolayer coverage determined by Karp et al.64 by the experimental standard heat of formation of atomic H in a gas phase (2.58 eV).65 

Figure 3 highlights the differences in our model, defined in (1) for (θ^H*,θ^O*), resulting from correlations between ΔEH and ΔEO. Consider

(25)
FIG. 3.

Contour plot of p(θ^H*,θ^O*) in log-scale where warmer colors represents higher densities. The upper contour plot with cooler colors corresponds to the uncorrelated case, which suggests that the density function in this case is flatter; the lower contour plot with warmer colors corresponds to the correlated case which suggests the density function has a higher mode located close to the bottom and the left of the figure. The model with correlations has significantly lower variance than the uncorrelated one, yielding an overall more predictive model.

FIG. 3.

Contour plot of p(θ^H*,θ^O*) in log-scale where warmer colors represents higher densities. The upper contour plot with cooler colors corresponds to the uncorrelated case, which suggests that the density function in this case is flatter; the lower contour plot with warmer colors corresponds to the correlated case which suggests the density function has a higher mode located close to the bottom and the left of the figure. The model with correlations has significantly lower variance than the uncorrelated one, yielding an overall more predictive model.

Close modal

where Π(θ^H*,θ^O*|ΔEH,ΔEO)=δ(θ^H*(ΔEH,ΔEO),θ^O*(ΔEH,ΔEO)), δ is the standard Dirac function and both θ^H*(ΔEH,ΔEO) and θ^O*(ΔEH,ΔEO) are given by (21).

Then, for the uncorrelated case (subscript uc), we can assume that

(26)

where pEH) and pEO) are defined as density functions of Gamma distributions, given by (30) and (31) in next section.

In the correlated case (subscript c), we assume that

(27)

where pEH) is still given by a Gamma distribution but pEOEH) comes from a normal distribution with mean aΔEH + b and variance determined by the data in Section VII B, which gives pcEH, ΔEO) a lower variance than pucEH, ΔEO).

We can use changing of variables such that

(28)

where J is the Jacobian of the inverse of coverage function θ^(ΔEH,ΔEO) from (8), evaluated at (θ^H*,θ^O*). Figure 3 shows the density function contours p(θ^H*,θ^O*), in the uncorrelated case using (26), and correlated case using (27). Note that the correlation of ΔEH and ΔEO reduces the variance of our model.

In this section, we compute the uncorrelated local and global sensitivity index defined in Section II B for the coverages θ^H* and θ^O* with respect to ΔEH, and will turn to the correlated cases in the next three sections. In section VI A, we compute the LSIs according to (21); and in section VI B, we construct an uncorrelated prior distribution for ΔEH and ΔEO, and then use this distribution and LSIs to compute GSIs.

Using the binding energies of adsorbed hydrogen and oxygen from Figure 2, we can analyze the relative LSIs for θ^H* and θ^O* with respect to ΔEH. SHH and SHO are identified using (5) and the model given in (21) (detailed calculations are presented in Section S2 of the supplementary material). For Pt, ΔEH = 2.6581(eV), ΔEO = 3.6604(eV), and the sensitivity of H and O coverages with respect to the H binding energy are SHH=38.9080 and SHO=0.0138. As expected, the H binding energy has a large effect on its coverage and a slight effect on the O coverage (some coupling is expected due to the competitive nature of adsorption).

To compute the corresponding GSIs on Pt by (7), we need to construct the distribution of our parameters, pEH, ΔEO). In the uncorrelated case, we have

(29)

where pEH) is the prior information for ΔEH on Pt, and pEOEH) = pEO) since ΔEH and ΔEO are independent, assuming no correlation.

Using the experimental and DFT data shown in Table I, we construct an informative prior for ΔEHEH. Let xH = 2.63 (eV), xO = 3.71 (eV), yH = 2.69 (eV) and yO = 3.68 (eV), where xi are the values given by experiment and yi are given by DFT, i = H, O. To quantify uncertainty from DFT error, we assume that ΔEH on Pt follows a gamma distribution with mean xH and the standard deviation given by the difference between experiment and DFT, (xiyi). We can construct the distribution for ΔEOEO in the same way under the uncorrelated assumption. The explicit density functions are shown below,

(30)
(31)

where ai=xi2/(xiyi)2 and bi=(xiyi)2/xi, i = H, O.

The GSIs, ξHH and ξHO, are formulated by (7) with q = 2 and p(λ) = pEH)pEO). Computing the integral in (7) numerically gives ξHH=1509.8865 and ξHO=0.2194. Again, the H binding energy has a major effect only on the H coverage and a slight effect on the O coverage.

The following sections cover the CLSI. In section VII A, we consider the simplest correlation model, both deterministic and linear, to compute the CLSIs defined in Section III. In section VII B, we construct parametric models for pEOEH) using the data shown in Figure 2, and compute the corresponding CLSIs. Section VIII covers non-parametric models.

To calculate the CLSI for θ^H* and θ^O* with respect to ΔEH, with the formula defined in (12), we use the conditional probability pEOEH). In the deterministic case, we can assume the conditional distribution of EO has a mean of gEH) and zero variance whose PDF can be described by the standard Dirac function

(32)

as presented in Section III. Then, using the data from Figure 2, one can determine the function gEH) with different fitting models, like polynomials and smoothing splines. In this paper, we use the linear function and set gEH) = a ΔEH + b, as shown in Figure 2. Then, the conditional distribution can be written as

(33)

For brevity, we only consider the CLSI formulation, as defined in (12) for the hydrogen coverage θ^H*, with respect to adsorbed hydrogen binding energy on the surface. The rest of CLSIs can be formulated by following the same procedure.

The CLSI for θ^H* with respect to ΔEH at the nominal hydrogen binding energy of ΔEH^ according to (12) takes the following form

(34)

where

(35)

The CLSI derivations in the presence of deterministic linear correlation are briefly described in Section S2 of the supplementary material and the results of SH,corrH and SH,corrO are shown in FIG. 1. The numerical results for Pt are SH,corrH=38.9021 and SH,corrO=97.7442. The corresponding uncorrelated LSI indices from (5) are SHH=38.9080, and SHO=0.0138; hence the correlation between ΔEH and ΔEO does not affect the sensitivity of θ^H* with respect to ΔEH, but does impact the sensitivity of θ^O* with respect to ΔEH. The LSI changes from slightly negative to highly positive. This is rationalized from the slope of the correlation depicted in Figure 2. Specifically, an increase in the binding energy of H leads a much higher increase in the binding energy of O and thus to an increase of the O coverage.

The above is a perfect linear model (deterministic) and ignores the variation around the linear fit of the binding energies. To capture correlations from the linear model, we set up a linear probabilistic model for ΔEH and ΔEO by introducing a random variable, ω, in the correlation,68 

(36)

To determine the distribution of ω, we can fit the data or adjusted data (to match the required domain of some distribution) using parametric models, like normal or gamma. Here we choose the normal distribution and fit the parameters using MATLAB by the Maximum Likelihood Estimation (MLE) method.68 The result is shown in Figure 4.

FIG. 4.

Model fitting for the random variable ω in (36) using a normal distribution; here we compare the best fit to the data’s histogram. The normal distribution is not a good approximation for the data since it does not properly capture the outlier values between -1 and -0.5, depicted in the histogram. Other parametric models give similar results (see Section S3 of the supplementary material).

FIG. 4.

Model fitting for the random variable ω in (36) using a normal distribution; here we compare the best fit to the data’s histogram. The normal distribution is not a good approximation for the data since it does not properly capture the outlier values between -1 and -0.5, depicted in the histogram. Other parametric models give similar results (see Section S3 of the supplementary material).

Close modal

Using (12) to compute the CLSI of θ^H* with respect to ΔEH, we consider

(37)

where p(ω) is the PDF for ω in (36). Instead of using the Monte Carlo method (discussed in Section S5 of the supplementary material), we can also use numerical integration to approximate the integral in (37).

The CLSI at the nominal point ΔEH* can then be obtained by direct differentiation

(38)

The gradient of lnFHH(ΔEH) is commonly estimated, such that

(39)

The CLSI for θ^O*, SH,corrO(ΔEH*), is computed similarly. The numerical results for Pt are SH,corrH=35.9874 and SH,corrO=9.4965. Compared to the deterministic model results in the previous subsection, we find that uncertainty significantly impacts SH,corrO. This is a rather interesting result because the correlation in the data (linear) results in the H binding energy having a significant effect on the O coverage but uncertainty significantly diminishes this effect. We give results from other parametric models in Section S3 of the supplementary material.

For small data sets, such as ours, parametric models are not usually adequate. Instead, we consider non-parametric methods.69 A possible non-parametric method is the empirical distribution function,

(40)

where I is the identity function. With this method, EP^[f] for some function f can be approximated via bootstrapping.69 

For categorical distributions, the bootstrap distribution is close to the posterior distribution with a non-informative symmetric Dirichlet prior according to Bayes method. It also has the same support, mean, and nearly the same covariance matrix as the data in the histogram. The bootstrap distribution is obtained without specifying either the prior or sampling from the posterior distribution.70 

We can also use curve estimation for our model.69 A simple density estimator is a histogram, which is a piece-wise constant function where the height of the function is proportional to number of observations in each bin

(41)

where B1, , Bn are the histogram bins, h = 1/n is the bin-width, and νi is the number of observations in Bi, as shown in Figure 4.

Smoother estimators, called kernel density estimators,69 converge faster to the true density than fitting from histograms because histograms are discontinuous

(42)

where h > 0 is the bandwidth and K is the kernel, defined to be any smooth function satisfying K(x) ≥ 0, ∫ K(x)dx = 1, ∫ xK(x)dx = 0 and σK2=x2K(x)dx>0.

In the main text of this work we use the histogram to approximate the distribution of ω, and use

(43)

to compute SH,corrH(ΔEH*) and SH,corrO(ΔEH*) using Equation 38. The numerical results for adsorption on Pt are SH,corrH=35.2196 and SH,corrO=12.4873. Results from the kernel density estimators with uniform and standard normal kernel, N(0,1), are given in Section S4 of the supplementary material.

Figure 5 summarizes all local sensitivity analysis results (the magnitude is plotted so a semi-log scale can be used). Correlations play a significant role as demonstrated in our earlier work.22 Clearly, the uncertainty in the correlations must properly by accounted for and, given the limited number of data we have for physical models, non-parametric models of the uncertainty are essential. For a large sample size, both (parametric and non-parametric) models should converge to the real distribution.36,69 Because we only have a few samples here, the non-parametric models approximate the noise term better, as shown in Figure 4.

FIG. 5.

Results of SHH, SHO and SH,corrH, SH,corrO for Pt. The bandwidth of histogram is 0.1. The sensitivities of θ^H* with respect to ΔEH are almost identical for uncorrelated and correlated models. However, the correlation between ΔEH and ΔEO significantly impacts the sensitivity of θ^O* on ΔEH, and changes the correlation from being slightly negative to highly positive. The overall shift in correlation is three orders of magnitude. Furthermore, the uncertainty ω in (36) also has a significant effect on SH,corrO: using a stochastic (parametric or non-parametric) model yields a sensitivity index smaller than the value from the deterministic model by an order of magnitude.

FIG. 5.

Results of SHH, SHO and SH,corrH, SH,corrO for Pt. The bandwidth of histogram is 0.1. The sensitivities of θ^H* with respect to ΔEH are almost identical for uncorrelated and correlated models. However, the correlation between ΔEH and ΔEO significantly impacts the sensitivity of θ^O* on ΔEH, and changes the correlation from being slightly negative to highly positive. The overall shift in correlation is three orders of magnitude. Furthermore, the uncertainty ω in (36) also has a significant effect on SH,corrO: using a stochastic (parametric or non-parametric) model yields a sensitivity index smaller than the value from the deterministic model by an order of magnitude.

Close modal

In this section we compute the CGSIs for the correlation model previously used to determine the CLSIs. According to (13), the CGSIs are

(44)
(45)

As discussed in Section VII, we assume the prior distribution of ΔEH on Pt satisfies

(46)

with a PDF of

(47)

using the data in Table I according to (30). Then, from SH,corrH and SH,corrO, we numerically calculate the CGSIs according to (44) and (45). The results are shown in Figure 6. Correlations have only a slight effect on the H coverage as we consider the H binding energy as an independent variable and the O binding energy as the dependent parameter.

FIG. 6.

Uncorrelated and correlated GSI results, ξHH, ξHO and ξH,coorH, ξH,coorO of Pt, computed by (7), (44) and (45). The correlation between ΔEH and ΔEO does not influence the sensitivity of θ^H* with respect to ΔEH. Correlations do, however, impact the sensitivity of θ^O* with respect to ΔEH. For ξH,coorO, we find that the CGSI from the purely data-driven non-parametric model are significantly higher than the that from the parametric (normal distribution) model.

FIG. 6.

Uncorrelated and correlated GSI results, ξHH, ξHO and ξH,coorH, ξH,coorO of Pt, computed by (7), (44) and (45). The correlation between ΔEH and ΔEO does not influence the sensitivity of θ^H* with respect to ΔEH. Correlations do, however, impact the sensitivity of θ^O* with respect to ΔEH. For ξH,coorO, we find that the CGSI from the purely data-driven non-parametric model are significantly higher than the that from the parametric (normal distribution) model.

Close modal

In Section III and Section S5 of the supplementary material we discuss the computation of the proposed correlated sensitivity indices using either direct numerical integration or Monte Carlo methods. Here, we briefly discuss the use of the Polynomial Chaos Expansion (PCE) method as an alternative to numerical integration (which is limited by the dimensions of the parameter space).71,72 Polynomial Chaos methods rely on expanding the model f(λ), defined in (2) in a series expansion, resulting in an approximation of the type

(48)

where d is the order of expansion approximation, ci are the expansion coefficients and P(i)(λ) are the polynomials forming the basis {P(0), , P(d)}. The chosen polynomials are orthogonal with respect to the probability measure of λ, i.e.

(49)

where δkl is the Kronecker delta function, p(λ) is the PDF of model parameters λ and Λ denotes the parameter space. Usually, λ = (λ1, , λn) is assumed to be independent. Carefully selecting the distribution (Gaussian, Gamma, etc) allows the corresponding basis to be given through the Askey scheme73 and can be implemented using the software DAKOTA.74 Using a previous approximation (48) allows calculation of the variance-based global sensitivity index (Sobol’s indices) directly and without extra cost;75 see also the implementations in Ref. 74. For instance, in the context of the applications discussed here, and in Ref. 76, the authors use PCE to analyze the problem of global sensitivity analysis for chemical processes, assuming uniformly distributed, uncorrelated parameters.

PCE can also be generalized to arbitrary distributions with the non-parametric models considered here. Such models include the use of histograms or kernel-based distributions. Indeed, in Ref. 77, the authors introduce a PCE with arbitrary probability measures, which can be either discrete, continuous, or discretized continuous. This form of PCE can also be specified either analytically (as probability density/cumulative distribution functions) or numerically (as various histograms or as raw data sets, like the ones arising in non-parametric methods). Only a few moments of the underlying distribution, and not on the specific functional form of the probability distribution functions, are required for these methods. Therefore, these methods do not apply to distributions which are not characterized by their moments, such as the lognormal.

We also carry out PCE for parameters λ which have correlated components. Indeed, in Ref. 78, Navarro et al. give us a way to instruct PCE for general multivariate distributions with correlated variables. In our case, the Sobol’s indices are not necessarily positive, and the contribution due to correlation can completely cancel the contribution from the variable itself, resulting in a small Sobol’s value even though such a variable can have a large impact on the outcome.78 It should be possible to apply the derivative-based sensitivity, as defined in section II B, by replacing f(λ) with the approximate PCE of the model. And it is also possible to combine the methods of Ref. 77 for the non-parametric aspects of the problem, and use Ref. 78 to address the correlations in the parameters. We expect to return to this implementation of PCE for non-parametric correlative sensitivity analysis in future work.

In this paper we proposed a non-parametric method for the local and global sensitivity analysis of models with correlated parameter dependencies. The resulting mathematical tools are applied on a benchmark Langmuir competitive adsorption model. Such systems are encountered in catalytic oxidation, such as emissions abatement, small scale power generation, fuel cells and batteries. In the system considered here, parameter correlations stem from correlated quantum-scale computational data. The necessity of using non-parametric methods arose from the limited amount of available quantum-scale data. In our methodology, we employed gradient-based methods to compute correlative local and global sensitivity indices to illustrate the relative effects of parameter perturbations (or errors and uncertainties) in the hydrogen and oxygen binding energies on the coverages. We observed that identification of influential parameters depends critically on whether or not correlations between parameters are taken into account. Furthermore, the impact of uncertainty in the correlation and the necessity of non-parametric approaches on the sensitivity indices are demonstrated. Finally, we briefly discussed the applicability of Polynomial Chaos expansion methods for the efficient simulation of sensitivity indices.

See supplementary material for the following topics relevant to this work: the detailed derivation of the Langmuir bimolecular adsorption model in Section S1; the derivation of LSIs for the model in Section S2; results of CLSIs with various parametric and non-parametric models in Section S3 and S4; finally, a computational implementation of CGSI by a Monte Carlo method in Section S5.

Financial support from the Defense Advanced Research Project Agency under grants R115-1342R1 and W911NF-15-2-0122 is gratefully acknowledged.

1.
C.
Goldsmith
,
A.
Tomlin
, and
S. J.
Klippenstein
,
Proc. Combust. Inst.
34
(
1
),
177
(
2013
).
2.
S.
Klippenstein
,
L.
Harding
,
M.
Davis
,
A.
Tomlin
, and
R.
Skodje
,
Proc. Combust. Inst.
33
(
1
),
351
(
2011
).
3.
O.
Le Maitre
and
O.
Knio
,
Spectral Methods for Uncertainty Quantification: With Applications to Computational Fluid Dynamics
(
Springer
,
New York, NY
,
2010
).
4.
J.
Murphy
,
D.
Sexton
,
D.
Barnett
,
G.
Jones
,
M.
Webb
,
M.
Collins
, and
D.
Stainforth
,
Nature
430
,
768
(
2004
).
5.
H.
Najm
,
Annual Review of Fluid Mechanics
41
,
135
(
2009
).
6.
J.
Prager
,
H.
Najm
, and
J.
Zador
,
Proc. Combust. Inst.
34
(
1
),
583
(
2013
).
7.
P.
Roache
,
Annual Review of Fluid Mechanics
29
,
123
(
1997
).
8.
R.
Smith
,
Uncertainty quantification: Theory, implementation, and applications
(
SIAM
,
Philadelphia, PA
,
2014
).
9.
N.
Vu-Bac
,
R.
Rafiee
,
X.
Zhuang
,
T.
Lahmer
, and
T.
Rabczuk
,
Composites Part B: Engineering
68
,
446
(
2015
).
10.
H.
Agarwal
,
J.
Renaud
,
E.
Preston
, and
D.
Padmanabhan
,
Reliability Engineering & System Safety
85
(
13
),
281
(
2004
).
11.
P.
Nissenson
,
J.
Thomas
,
B.
Finlayson-Pitts
, and
D.
Dabdub
,
Atmospheric Environment
42
(
29
),
6934
(
2008
).
12.
M.
Reagan
,
H.
Najm
,
R.
Ghanem
, and
O.
Knio
,
Combustion and Flame
132
(
3
),
545
(
2003
).
13.
P.
Dupuis
,
M. A.
Katsoulakis
,
Y.
Pantazis
, and
P.
Plecháč
,
SIAM/ASA Journal on Uncertainty Quantification
4
,
80
(
2016
).
14.
M.
Haaker
and
P.
Verheijen
,
Chemical Engineering Research and Design
183
(
5
),
591
(
2004
).
15.
J.
Zador
,
I.
Zsely
, and
T.
Turanyi
,
Reliability Engineering & System Safety
91
(
10-11
),
1232
(
2006
).
16.
S.
Da Veiga
,
F.
Wahl
, and
F.
Gamboa
,
Technometrics
51
(
4
),
452
(
2009
).
17.
S.
Kucherenko
,
S.
Tarantola
, and
P.
Annoni
,
Comput. Phys. Commun.
183
(
4
),
937
(
2012
).
18.
P.
Li
and
Q.
Vu
,
BMC Systems Biology
7
(
1
),
1
(
2013
).
19.
T.
Mara
and
T.
Stefano
,
Reliability Engineering & System Safety
107
,
115
(
2012
).
20.
T.
Nagy
and
T.
Turányi
,
Reliability Engineering & System Safety
107
,
29
(
2012
).
21.
R.
Shannon
,
A.
Tomlin
,
S.
Robertson
,
M.
Blitz
,
M.
Pilling
, and
P.
Seakins
,
J. Phys. Chem. A
119
(
28
),
7430
(
2015
).
22.
J.
Sutton
,
W.
Guo
,
M.
Katsoulakis
, and
D.
Vlachos
,
Nature Chemistry
(
2016
).
23.
A.
Tomlin
,
E.
Agbro
,
V.
Nevrly
,
J.
Dlabka
, and
M.
Vasinek
,
Int. J. Chem. Kinet.
46
(
11
),
662
(
2014
).
24.
T.
Ziehn
,
K.
Hughes
,
J.
Griffiths
,
R.
Porter
, and
A.
Tomlin
,
Combustion Theory and Modelling
13
(
4
),
589
(
2009
).
25.
T.
Ziehn
and
A.
Tomlin
,
Int. J. Chem. Kinet.
40
(
11
),
742
(
2008
).
26.
T.
Ziehn
and
A.
Tomlin
,
Environmental Modelling & Software
24
(
7
),
775
(
2009
).
27.
I.
Zsely
,
J.
Zador
, and
T.
Turanyi
,
Int. J. Chem. Kinet.
40
(
11
),
754
(
2008
).
28.
G. N.
Derry
and
P. N.
Ross
,
The Journal of Chemical Physics
82
,
2772
(
1985
).
29.
H. A.
Engelhardt
and
D.
Menzel
,
Surface Science
57
,
591
(
1976
).
30.
T. W.
Root
,
L. D.
Schmidt
, and
G. B.
Fisher
,
Surface Science
134
,
30
(
1983
).
31.
G.
Blyholder
,
The Journal of Physical Chemistry
68
,
2772
(
1964
).
32.
L.
Loulergue
,
A.
Schilt
,
R.
Spahni
,
V.
Masson-Delmotte
,
T.
Blunier
,
B.
Lemieux
,
J.-M.
Barnola
,
D.
Raynaud
,
T. F.
Stocker
, and
J.
Chappellaz
,
Nature
453
,
383
(
2008
).
34.
A. K.
Tripati
,
C. D.
Roberts
, and
R. A.
Eagle
,
Science
326
,
1394
(
2009
).
35.
A.
Raftery
,
D.
Madigan
, and
J.
Hoeting
,
Journal of the American Statistical Association
92
,
179
(
1997
).
36.
L.
Wasserman
,
All of Statistics: A Concise Course in Statistical Inference
(
Springer
,
New York, NY
,
2004
).
37.
A.
Gelman
,
J.
Carlin
,
H.
Stern
,
D.
Dunson
,
A.
Vehtari
, and
D.
Rubin
,
Bayesian Data Analysis
(
CRC Press
,
Boca Raton, FL
,
2014
).
38.
A.
Ayeb
,
W.
Otten
,
A.
Mankb
, and
P.
Notten
,
Journal of the Electrochemical Society
153
(
11
),
A2055
(
2006
).
39.
S.
Gu
,
B.
Xu
, and
Y.
Yan
,
Annual Review of Chemical and Biomolecular Engineering
5
,
429
(
2014
).
40.
S.
Lee
,
S.
Mukerjee
,
E.
Ticianelli
, and
J.
McBreen
,
Electrochimica Acta
44
(
19
),
3283
(
1999
).
41.
S.
Liu
,
T.
Ishimoto
,
D.
Monder
, and
M.
Koyama
,
The Journal of Physical Chemistry C
119
(
49
),
27603
(
2015
).
42.
T.
Nagasawa
and
K.
Hanamura
,
Journal of Power Sources
290
,
168
(
2015
).
43.
I.
Chorkendorff
and
J.
Niemantsverdriet
,
Concepts of Modern Catalysis and Kinetics
(
Wiley
,
2006
).
44.
M.
Davis
and
R.
Davis
,
Fundamentals of chemical reaction engineering
(
McGraw-Hill
,
New York, NY
,
2003
).
45.
R.
Masel
,
Chemical kinetics and catalysis
(
Wiley
,
New York, NY
,
2001
).
46.
A.
Groß
and
A.
Dianat
,
Physical Review Letters
98
,
206107
(
2007
).
47.
A.
Groß
,
The Journal of Chemical Physics
135
,
174707
(
2011
).
48.
K.
Sillar
,
A.
Kundu
, and
J.
Sauer
,
The Journal of Physical Chemistry C
121
,
12789
(
2017
).
49.
R.
Carr
,
J.
Comer
,
M. D.
Ginsberg
, and
A.
Aksimentiev
,
The Journal of Physical Chemistry Letters
2
,
1804
(
2011
).
50.
G.
Kresse
and
J.
Hafner
,
Physical Review B
47
,
558
(
1993
).
51.
G.
Kresse
and
J.
Hafner
,
Physical Review B
49
,
14251
(
1994
).
52.
G.
Kresse
and
J.
Furthmller
,
Computational Materials Science
6
,
15
(
1996
).
53.
G.
Kresse
and
J.
Furthmller
,
Physical Review B
54
,
11169
(
1996
).
54.
P. E.
Blöchl
,
Physical Review B
50
,
17953
(
1994
).
55.
G.
Kresse
and
D.
Joubert
,
Physical Review B
59
,
1758
(
1999
).
56.
A. V.
Mironenko
,
M. J.
Gilkey
,
P.
Panagiotopoulou
,
G.
Facas
,
D. G.
Vlachos
, and
B.
Xu
,
The Journal of Physical Chemistry C
119
,
6075
(
2015
).
57.
S. R.
Bahn
and
K. W.
Jacobsen
,
Computing in Science and Engineering
4
,
56
(
2002
).
58.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Physical Review Letters
77
,
3865
(
1996
).
59.
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
H.
Krieg
,
The Journal of Chemical Physics
132
,
154104
(
2010
).
60.
J.
Monkhurst
and
J. J.
Pack
,
Physical Review B, Solid State
13
,
5188
(
1976
).
61.
U.
Fink
,
T. A.
Wiggins
, and
D. H.
Rank
,
Journal of Molecular Spectroscopy
18
,
384
(
1965
).
62.
A.
Weber
and
E. A.
McGinnis
,
Journal of Molecular Spectroscopy
4
,
195
(
1960
).
63.
D.
McQuarrie
,
Statistical Mechanics
(
University Science Books
,
2000
).
64.
E. M.
Karp
,
C. T.
Campbell
,
F.
Studt
,
F.
Abild-Pedersen
, and
J. K.
Nrskov
,
The Journal of Physical Chemistry C
116
,
25772
(
2012
).
65.
M.
Chase
, Jr.
, “
Nist-janf thermochemical tables
,” (
2016
), accessed: 2016-05-31.
66.
B. d.
Darwent
,
Natl. Stand. Ref. Data Ser. (U.S., Natl. Bur. Stand.)
31
(
32
),
41
(
1970
).
67.
B.
Hammer
,
Physical Review B
59
,
7413
(
1999
).
68.
G.
James
,
D.
Witten
,
T.
Hastie
, and
R.
Tibshirani
,
An introduction to statistical learning
, Vol. 6 (
Springer
,
2013
).
69.
L.
Wasserman
,
All of Nonparametric Statistics
(
Springer
,
New York, NY
,
2006
).
70.
J.
Friedman
,
T.
Hastie
, and
R.
Tibshirani
,
The elements of statistical learning
, Chapter 8, Vol. 1 (Springer series in statistics
Springer
,
Berlin
,
2001
).
71.
R. G.
Ghanem
and
P. D.
Spanos
,
Stochastic finite elements: a spectral approach
(
Courier Corporation
,
2003
).
72.
D.
Xiu
,
Numerical methods for stochastic computations: a spectral method approach
(
Princeton university press
,
2010
).
73.
D.
Xiu
and
G. E.
Karniadakis
,
SIAM Journal on Scientific Computing
24
,
619
(
2002
).
74.
B. M.
Adams
,
M. S.
Ebeida
,
M. S.
Eldred
,
J. D.
Jakeman
,
L. P.
Swiler
,
J. A.
Stephens
,
D. M.
Vigil
,
T. M.
Wildey
,
W. J.
Bohnhoff
,
J. P.
Eddy
 et al., “
Dakota, a multilevel parallel object-oriented framework for design optimization, parameter estimation, uncertainty quantification, and sensitivity analysis
,” Tech. Rep. (
Sandia National Laboratories (SNL-NM)
,
Albuquerque, NM (United States
),
2014
).
75.
T.
Crestaux
,
O.
Le Maıtre
, and
J.-M.
Martinez
,
Reliability Engineering & System Safety
94
,
1161
(
2009
).
76.
P. L. T.
Duong
,
W.
Ali
,
E.
Kwok
, and
M.
Lee
,
Computers & Chemical Engineering
90
,
23
(
2016
).
77.
S.
Oladyshkin
and
W.
Nowak
,
Reliability Engineering & System Safety
106
,
179
(
2012
).
78.
M.
Navarro
,
J.
Witteveen
, and
J.
Blom
, arXiv preprint arXiv:1406.5483 (
2014
).

Supplementary Material