11institutetext: Francesco Dell’Accio 22institutetext: Department of Mathematics and Computer Science, University of Calabria, Rende (CS), Italy
22email: [email protected]
33institutetext: Mohammed Kbiri 44institutetext: Department of Mathematics, College of Science, King Khalid University, P.O. Box 9004, 61413 Abha, Saudi Arabia.
44email: [email protected]
55institutetext: Allal Guessab 66institutetext: Avenue Al Walae, 73000, Dakhla, Morocco
66email: [email protected]
77institutetext: Federico Nudo (corresponding author) 88institutetext: Department of Mathematics and Computer Science, University of Calabria, Rende (CS), Italy
88email: [email protected]

A General Probability Density Framework for Local Histopolation and Weighted Function Reconstruction from Mesh Line Integrals

Francesco Dell’Accio    Allal Guessab    Mohammed Kbiri Alaoui    Federico Nudo
(Version: September 03, 2025)
Abstract

In this paper, we study the reconstruction of a bivariate function from weighted integrals along the edges of a triangular mesh, a problem of central importance in tomography, computer vision, and numerical approximation. Our approach relies on local histopolation methods defined through unisolvent triples, where the edge weights are induced by suitable probability densities. In particular, we introduce two new two-parameter families of generalized truncated normal distributions, which extend classical exponential-type laws and provide additional flexibility in capturing local features of the target function. These distributions give rise to new quadratic reconstruction operators that generalize the standard linear histopolation scheme, while retaining its simplicity and locality. We establish their theoretical foundations, proving unisolvency and deriving explicit basis functions, and we demonstrate their improved accuracy through extensive numerical tests. Moreover, we design an algorithm for the optimal selection of the distribution parameters, ensuring robustness and adaptivity of the reconstruction. Finally, we show that the proposed framework naturally extends to any bivariate function whose restriction to the edges defines a valid probability density, thus highlighting its generality and broad applicability.

1 Introduction

The classical concept of polynomial interpolation can be extended beyond pointwise evaluations to situations where the available information consists of general linear functionals Rivlin. When these data take the form of integrals over geometric domains, the problem is referred to as histopolation Robidoux, and the corresponding techniques are known as histopolation methods.

Histopolation provides a general framework for reconstructing functions from integral data, a problem of central importance in many areas of science and engineering. In practical applications, direct pointwise samples of the target function are often unavailable; instead, only integral measurements are accessible. This situation arises prominently in computed tomography and medical imaging, where measurements typically consist of line integrals kak2001principles; natterer2001mathematics; palamodov2016reconstruction. Beyond tomography, histopolation has found applications in computer vision, signal and image processing Bosner:2020:AOC, curve and spline approximation Fischer:2005:MPR; Fischer:2007:CSP; Siewer:2008:HIS; Hallik:2017:QLR, fractal functions Barnsley:2023:HFF, and in the conservation of physical quantities in numerical simulations HiemstraJCP. Recently, histopolation methods have received increasing attention BruniErbFekete; bruni2024polynomial, including global polynomial histopolation–regression in one bruni2025polynomial and several dimensions bruno2025bivariate, as well as weighted local polynomial histopolation schemes in the bivariate setting nudo1; nudo2; dell2025reconstructing; dell2025truncated; milov1.

A standard strategy for these reconstruction problems is to partition the computational domain into subdomains (a mesh), approximate locally on each element, and assemble a global approximation from the local contributions. Such schemes are typically conforming or nonconforming, depending on whether the assembled approximation is continuous across element interfaces.

A local approximation method can be rigorously defined through the triple Guessab:2022:SAB

d=(Sd,𝔽Sd,ΣSd),\mathcal{M}_{d}=\left(S_{d},\mathbb{F}_{S_{d}},\Sigma_{S_{d}}\right),

where

  • SdS_{d} is a polytope in d\mathbb{R}^{d}, d1d\geq 1;

  • 𝔽Sd\mathbb{F}_{S_{d}} is an nn-dimensional space of trial functions on SdS_{d};

  • ΣSd={j:j=1,,n}\Sigma_{S_{d}}=\left\{\mathcal{L}_{j}:j=1,\dots,n\right\} is a set of linearly independent linear functionals, called the degrees of freedom.

The triple d\mathcal{M}_{d} is said to be unisolvent if, whenever f𝔽Sdf\in\mathbb{F}_{S_{d}} satisfies

j(f)=0,j=1,,n,\mathcal{L}_{j}(f)=0,\quad j=1,\dots,n,

then necessarily f=0f=0. A set of functions

B={φi:i=1,,n}B=\left\{\varphi_{i}\,:\,i=1,\dots,n\right\}

such that

span{φi:i=1,,n}=Sd\operatorname{span}\left\{\varphi_{i}\,:\,i=1,\dots,n\right\}=S_{d}

and

j(φi)=δij\mathcal{L}_{j}\left(\varphi_{i}\right)=\delta_{ij}

is said to form the basis functions associated with the triple d\mathcal{M}_{d}. If the degrees of freedom are integral functionals, the scheme is also referred to as a local histopolation method.

The simplest example is the linear nonconforming histopolation method on triangles, where edge integrals act as degrees of freedom and piecewise linear functions form the trial space. While attractive for its simplicity and efficiency, this approach has important limitations: its piecewise linear structure fails to reproduce oscillatory behaviors or sharp gradients, and higher accuracy can only be achieved through substantial mesh refinement.

In this work we propose an enrichment strategy that enhances local histopolation schemes by introducing quadratic trial spaces together with weighted degrees of freedom. The key idea is to employ probability densities on the edges of the mesh, which act as flexible weights and introduce tunable parameters. In particular, we construct two new two-parameter families of generalized truncated normal distributions, which extend classical exponential-type laws and include special or limiting cases (such as truncated normal and beta-type densities). These families naturally lead to enriched quadratic operators that significantly improve reconstruction accuracy while retaining the locality and simplicity of the original method.

From a theoretical point of view, we prove unisolvency of the enriched triples and derive explicit closed-form basis functions. From a practical perspective, we validate the proposed operators through extensive numerical experiments, which demonstrate their effectiveness in reconstructing both smooth and non-smooth functions. Moreover, as shown in Section 3, the framework is not limited to the two families introduced here: the same reasoning applies to any bivariate function whose restriction to the edges defines a valid probability density, thus highlighting the generality and robustness of the proposed approach.

The paper is organized as follows. In Section 2, we introduce two families of generalized truncated normal distributions, analyze their main properties, and define the corresponding families of quadratic histopolation operators. Moreover, we describe an algorithm for the optimal selection of the distribution parameters, which further enhances the robustness and adaptivity of the reconstruction. In Section 3, we develop the general framework for enriched quadratic histopolation operators and establish their theoretical foundations. In Section 4, we present numerical experiments that confirm the accuracy and effectiveness of the proposed methods. Finally, in Section 5, we conclude with a summary of the contributions and discuss possible directions for future research, including three-dimensional extensions and applications to imaging and approximation problems.

2 Two-parameter families of generalized truncated normal distributions

2.1 Problem statement

The problem we want to address is the following: given only integral measurements of an unknown bivariate function along the edges of a triangular mesh, reconstruct an accurate approximation of the function inside each triangle. In the classical local histopolation method, these measurements are taken as unweighted edge integrals and the reconstruction space is piecewise linear. While simple and efficient, this approach lacks the ability to capture curved or oscillatory features unless the mesh is strongly refined.

Our objective is to overcome this limitation by enriching the standard framework. We introduce probability density functions as edge weights in the integral data, so that each measurement carries additional information about the local behavior of the function. This enrichment is feasible under the assumption that additional integral data are available, beyond the classical unweighted edge averages. In this setting, the reconstruction naturally leads to quadratic operators that preserve locality and simplicity while significantly improving approximation accuracy. The role of this section is to present two flexible two-parameter distribution families that serve as the building blocks of the proposed weighted histopolation scheme.

2.2 Definition of the distribution families

In this section, we introduce two distinct two-parameter families of univariate distributions, each generated by suitable bivariate weight functions defined on a triangle of a given mesh. These families include, as limiting or special cases, several well-known probability laws. They provide the building blocks for a broad class of new piecewise quadratic reconstruction operators, which in turn form the basis of a nonconforming histopolation method locally determined by a unisolvent triple. The goal is to reconstruct a function using information derived from weighted integrals over the edges of a triangular mesh. To this end, we formulate the probability density functions (hereafter, pdfs) associated with the edges of a nondegenerate triangle TT with vertices 𝒗1,𝒗2,𝒗3\boldsymbol{v}_{1},\boldsymbol{v}_{2},\boldsymbol{v}_{3} and barycentric coordinates λ1,λ2,λ3\lambda_{1},\lambda_{2},\lambda_{3}. These coordinates are the unique affine functions on TT, namely

λi(t𝒙+(1t)𝒚)=tλi(𝒙)+(1t)λi(𝒚),𝒙,𝒚T,t[0,1],\lambda_{i}\left(t\boldsymbol{x}+(1-t)\boldsymbol{y}\right)=t\lambda_{i}(\boldsymbol{x})+(1-t)\lambda_{i}(\boldsymbol{y}),\quad\boldsymbol{x},\boldsymbol{y}\in T,\quad t\in[0,1], (1)

which satisfy the partition of unity property,

i=13λi(𝒙)=1,𝒙T,\sum_{i=1}^{3}\lambda_{i}(\boldsymbol{x})=1,\quad\boldsymbol{x}\in T,

and the Kronecker delta property at the vertices,

λi(𝒗j)=δij={1,i=j,0,ij,i,j=1,2,3.\lambda_{i}\left(\boldsymbol{v}_{j}\right)=\delta_{ij}=\begin{cases}1,&i=j,\\[4.0pt] 0,&i\neq j,\end{cases}\quad i,j=1,2,3.

As a consequence, we have

λi(𝒙)=0,𝒙si,i=1,2,3.\lambda_{i}(\boldsymbol{x})=0,\quad\boldsymbol{x}\in s_{i},\quad i=1,2,3. (2)

For the forthcoming analysis, we recall the lower incomplete gamma function, which will play a central role. It is defined for s>0s>0 and z>0z>0 by

γ(s,z)=0zts1et𝑑t.\gamma(s,z)=\int_{0}^{z}t^{s-1}e^{-t}dt. (3)

Among its numerous properties, one that we shall exploit is the limiting relation

limz0γ(s,z)zs=1s.\lim_{z\to 0}\frac{\gamma(s,z)}{z^{s}}=\frac{1}{s}. (4)

Indeed, since

ezet1,0tz,e^{-z}\leq e^{-t}\leq 1,\quad 0\leq t\leq z,

we obtain

ts1ezts1etts1.t^{s-1}e^{-z}\leq t^{s-1}e^{-t}\leq t^{s-1}.

Integrating over [0,z][0,z] with respect to tt and dividing by zsz^{s}, we obtain the inequalities

ezsγ(s,z)zs1s.\frac{e^{-z}}{s}\leq\frac{\gamma(s,z)}{z^{s}}\leq\frac{1}{s}.

Taking the limit as z0z\to 0 establishes (4). This relation will be crucial for analyzing the convergence of our densities when one of their parameters tends to infinity. For convenience, we introduce the modified incomplete gamma function

γmod(s,z):=γ(s,z)zs=1zs0zts1et𝑑t,\gamma^{\mathrm{mod}}(s,z):=\frac{\gamma(s,z)}{z^{s}}=\frac{1}{z^{s}}\int_{0}^{z}t^{s-1}e^{-t}dt, (5)

so that (4) takes the simpler form

limz0γmod(s,z)=1s.\lim_{z\to 0}\gamma^{\mathrm{mod}}(s,z)=\frac{1}{s}. (6)

If a numerical routine is available to evaluate γ(s,z)\gamma(s,z), then γmod(s,z)\gamma^{\mathrm{mod}}(s,z) can be computed directly by scaling. For example, in Matlab one may use gammainc(z,s) gamma(s), while in Mathematica the corresponding function is Gamma[s,0,z].

2.3 First family of distributions

The first family is defined through the following bivariate function with shape parameter μ1\mu\geq 1 and scale parameter σ>0\sigma>0 by

Kσ,μ(𝒙)=aσ,μ(H2(𝒙))μ1e12(H(𝒙)σ2)μ,𝒙T,K_{\sigma,\mu}(\boldsymbol{x})=a_{\sigma,\mu}\left(H^{2}(\boldsymbol{x})\right)^{\mu-1}e^{-\frac{1}{2}\left(\frac{H(\boldsymbol{x})}{\sigma^{2}}\right)^{\mu}},\quad\boldsymbol{x}\in T, (7)

where HH is expressed in terms of barycentric coordinates as

H(𝒙)=2i=13λi2(𝒙)1,𝒙T,H(\boldsymbol{x})=2\sum_{i=1}^{3}\lambda_{i}^{2}(\boldsymbol{x})-1,\quad\boldsymbol{x}\in T, (8)

and aσ,μa_{\sigma,\mu} is a normalization factor ensuring that integrals involving the associated generalized truncated normal distribution are properly scaled. It is worth noting that Kσ,μK_{\sigma,\mu} is a weight function, as it is nonnegative throughout the triangle TT. Furthermore, when restricted to any edge of TT, HH reduces to a quadratic form, and the induced univariate functions not only serve as weights but also define valid probability density functions on the edges. An advantage of the general form (7) is that it allows us to tune the free parameters σ\sigma and μ\mu when reconstructing a function with specific features. In particular, different behaviors can be produced by varying the exponential term in (7).

In the following, we parametrize the edge sjs_{j} as

1+t2𝒗j+1+1t2𝒗j+2,t[1,1].\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2},\quad t\in[-1,1]. (9)

Then, by (1) and (2), together with (8), we obtain

H(1+t2𝒗j+1+1t2𝒗j+2)=t2.H\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)=t^{2}. (10)

We now define the first class of univariate pdfs associated with the weight function Kσ,μK_{\sigma,\mu}. It follows directly from the definition of Kσ,μK_{\sigma,\mu} that on the edge sis_{i} of TT one has

kσ,μ(t):=Kσ,μ(1+t2𝒗j+1+1t2𝒗j+2)=aσ,μ(t2)2μ2e12(t2σ2)μ,t[1,1].k_{\sigma,\mu}(t):=K_{\sigma,\mu}\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)=a_{\sigma,\mu}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}},\quad t\in\left[-1,1\right]. (11)

An important observation is that kσ,μ(t)k_{\sigma,\mu}(t) is independent of the choice of edge of the triangle TT. We shall henceforth work with the normalized probability density functions (pdfs) with unit zeroth moment, which we denote with a tilde. Thus, in order to obtain a normalized pdf, the coefficient aσ,μa_{\sigma,\mu} in the bivariate weight function Kσ,μ(𝒙)K_{\sigma,\mu}(\boldsymbol{x}) defined in (7) is introduced so that

11kσ,μ(t)𝑑t=1.\int_{-1}^{1}k_{\sigma,\mu}(t)dt=1.

As shown in Lemma 2, for any μ1\mu\geq 1 and σ>0\sigma>0, the normalization constant aσ,μa_{\sigma,\mu} is given by

aσ,μ=μ24μ32μσ4μ3γ(4μ32μ,12σ2μ),a_{\sigma,\mu}=\frac{\mu}{2^{\frac{4\mu-3}{2\mu}}\sigma^{4\mu-3}\gamma\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)},

which can be written more compactly in terms of the modified incomplete gamma function as

aσ,μ=μγmod(4μ32μ,12σ2μ).a_{\sigma,\mu}=\frac{\mu}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}. (12)

Hence, for any σ>0\sigma>0 and μ1\mu\geq 1, substituting (12) into (11), it follows that the normalized pdfs on [1,1][-1,1] are given by

k~σ,μ(t)=μγmod(4μ32μ,12σ2μ)(t2)2μ2e12(t2σ2)μ,t[1,1].\widetilde{k}_{\sigma,\mu}(t)=\frac{\mu}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}},\quad t\in\left[-1,1\right]. (13)

As a first observation, this family extends the class of exponential-type distributions. Indeed, when μ=1\mu=1, equation (13) reduces to the well-known doubly truncated normal distribution on the interval [1,1][-1,1], here written in terms of γmod\gamma^{\mathrm{mod}}; see Jawitz2004. Indeed, by straightforward calculation, and using the notation (6), this distribution can be written as

k~σ,1(t)=1γmod(12,12σ2)e12(t2σ2),t[1,1].\widetilde{k}_{\sigma,1}(t)=\frac{1}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{2}}\right)}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)},\quad t\in\left[-1,1\right].

Recall that this latter distribution has no shape parameter, and thus cannot model all phenomena. Nevertheless, it has been extensively studied and plays a central role in probability theory, see, e.g., Xinyi2012functions and the references therein. Therefore, our pdfs constitute a new two-parameter family of exponential type. In Figure 1, we plot several instances of k~σ,μ\widetilde{k}_{\sigma,\mu}, illustrating the strong sensitivity of the distribution to both the shape parameter μ\mu and the scale parameter σ\sigma. This highlights the increased flexibility and importance of the generalized family.

Refer to caption
Refer to caption
Figure 1: Normalized probability density functions k~σ,μ\widetilde{k}_{\sigma,\mu} on [1,1][-1,1]. Left: effect of the scale parameter, with fixed μ=2\mu=2 and different values of σ\sigma. Right: effect of the shape parameter, with fixed σ=0.50\sigma=0.50 and varying values of μ\mu, illustrating the increased sharpness and concentration around t=0t=0. These plots emphasize the flexibility provided by the two parameters.
Remark 1

The pdfs k~σ,μ\widetilde{k}_{\sigma,\mu} also include a beta-type distribution as a limiting case. Indeed, using (6) with

s=4μ32μ,z=12σ2μ,s=\frac{4\mu-3}{2\mu},\quad z=\frac{1}{2\sigma^{2\mu}},

we obtain

limσγmod(4μ32μ,12σ2μ)=2μ4μ3.\lim_{\sigma\to\infty}\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)=\frac{2\mu}{4\mu-3}. (14)

Therefore, for any t[1,1]t\in[-1,1], the limiting distribution is

k~,μ(t):=limσk~σ,μ(t)=4μ32(t2)2μ2.\widetilde{k}_{\infty,\mu}(t):=\lim_{\sigma\to\infty}\widetilde{k}_{\sigma,\mu}(t)=\frac{4\mu-3}{2}\left(t^{2}\right)^{2\mu-2}. (15)

Thus, k~σ,μ\widetilde{k}_{\sigma,\mu} converges to a beta-type distribution as σ\sigma\to\infty.

In order to present our local reconstruction operator, which is based on the pdfs k~σ,μ\widetilde{k}_{\sigma,\mu}, we now introduce some additional notations and recall the basic setting for clarity of exposition. The operator is built locally on each nondegenerate triangle TT of a given mesh. For notational convenience, we set

𝒗4=𝒗1,𝒗5=𝒗2,\boldsymbol{v}_{4}=\boldsymbol{v}_{1},\quad\boldsymbol{v}_{5}=\boldsymbol{v}_{2},

as well as

λ4=λ1,λ5=λ2.\lambda_{4}=\lambda_{1},\quad\lambda_{5}=\lambda_{2}.

Then, using the parametrization of the side sjs_{j} given in (9), we introduce the linear functionals

jCH(f)=1|sj|sjf(𝒙)𝑑𝒙=1211f(1+t2𝒗j+1+1t2𝒗j+2)𝑑𝒙,j=1,2,3,\mathcal{I}_{j}^{\mathrm{CH}}(f)=\frac{1}{\left\lvert s_{j}\right\rvert}\int_{s_{j}}f(\boldsymbol{x})d\boldsymbol{x}=\frac{1}{2}\int_{-1}^{1}f\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)d\boldsymbol{x},\quad j=1,2,3, (16)

and denote

ΣCH={jCH:j=1,2,3}.\Sigma^{\mathrm{CH}}=\left\{\mathcal{I}_{j}^{\mathrm{CH}}\,:\,j=1,2,3\right\}.

Thus, the linear nonconforming histopolation method can be locally represented by the triple

𝒞=(T,1(T),ΣCH),\mathcal{CH}=\left(T,\mathbb{P}_{1}(T),\Sigma^{\mathrm{CH}}\right), (17)

where 1(T)\mathbb{P}_{1}(T) is the space of linear polynomials on TT, and its associated basis functions are expressed in barycentric coordinates as

φiCH=12λi,i=1,2,3,\varphi_{i}^{\mathrm{CH}}=1-2\lambda_{i},\quad i=1,2,3,

as can be directly verified. The reconstruction operator associated to the triple (17) is

πCH:C(T)1(T)fj=13jCH(f)φjCH.\begin{array}[]{rcl}{\pi}^{{\mathrm{CH}}}:C(T)&\rightarrow&{\mathbb{P}}_{1}(T)\\ f&\mapsto&\displaystyle{\sum_{j=1}^{3}}\mathcal{I}^{\mathrm{CH}}_{j}\left(f\right)\varphi^{\mathrm{CH}}_{j}.\end{array} (18)

To describe our first new two-parameter family of nonconforming histopolation operators, we introduce a class of enriched linear functionals, defined by

jenr(f)\displaystyle{\mathcal{I}}_{j}^{\mathrm{enr}}(f) =\displaystyle= 11f(1+t2𝒗j+1+1t2𝒗j+2)k~σ,μ(t)𝑑t,j=1,2,3,\displaystyle\int_{-1}^{1}f\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{k}_{\sigma,\mu}(t)dt,\quad j=1,2,3, (19)
jenr(f)\displaystyle{\mathcal{L}}_{j}^{\mathrm{enr}}(f) =\displaystyle= 11pσ,μ(t)f(1+t2𝒗j+1+1t2𝒗j+2)k~σ,μ(t)𝑑t,j=1,2,3,\displaystyle\int_{-1}^{1}{p}_{\sigma,\mu}(t)f\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{k}_{\sigma,\mu}(t)dt,\quad j=1,2,3, (20)

where

pσ,μ(t)=t2γmod(4μ12μ,12σ2μ)γmod(4μ32μ,12σ2μ).{p}_{\sigma,\mu}(t)=t^{2}-\frac{\gamma^{\mathrm{mod}}\left(\frac{4\mu-1}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}. (21)

Since k~σ,μ\widetilde{k}_{\sigma,\mu} is a probability density, the functionals jenr\mathcal{I}_{j}^{\mathrm{enr}} and jenr\mathcal{L}_{j}^{\mathrm{enr}} can be interpreted as weighted averages of ff and pσ,μfp_{\sigma,\mu}f along the side sjs_{j} of TT. The rationale behind the choice of pσ,μp_{\sigma,\mu}, which will become clear in Section 3, lies in its central role in the subsequent analysis. Using (10) and the parametrization (9), this polynomial can equivalently be expressed in terms of the function HH defined in (8) as

pσ,μ(t)\displaystyle{p}_{\sigma,\mu}(t) =\displaystyle= H~(1+t2𝒗j+1+1t2𝒗j+2)\displaystyle\widetilde{H}\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)
:=\displaystyle:= H(1+t2𝒗j+1+1t2𝒗j+2)γmod(4μ12μ,12σ2μ)γmod(4μ32μ,12σ2μ).\displaystyle H\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)-\frac{\gamma^{\mathrm{mod}}\left(\frac{4\mu-1}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}.
Remark 2

We remark that, using (11), the product pσ,μ(t)k~σ,μ(t)p_{\sigma,\mu}(t)\widetilde{k}_{\sigma,\mu}(t) appearing in the definition of jenr\mathcal{L}_{j}^{\mathrm{enr}} can be further written as

pσ,μ(t)k~σ,μ(t)=(H~Kσ,μ)(1+t2𝒗j+1+1t2𝒗j+2).p_{\sigma,\mu}(t)\widetilde{k}_{\sigma,\mu}(t)=\left(\widetilde{H}K_{\sigma,\mu}\right)\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right).

Moreover, we note that when μ=1\mu=1, equation (15) implies

k~,1(t)=12,\widetilde{k}_{\infty,1}(t)=\frac{1}{2},

and in this particular case the enriched functionals in (19) reduce to the classical unweighted histopolation functionals of (16).

We define the associated set of enriched linear functionals as

Σ1,σ,μenr(T)={jenr,jenr:j=1,2,3}\Sigma^{\mathrm{enr}}_{1,\sigma,\mu}(T)=\left\{{\mathcal{I}}_{j}^{\mathrm{enr}},{\mathcal{L}}^{\mathrm{enr}}_{j}\,:\,j=1,2,3\right\}

and introduce the enriched triple

1,σ,μenr=(T,2(T),Σ1,σ,μenr(T)),\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu}=\left(T,\mathbb{P}_{2}(T),\Sigma^{\mathrm{enr}}_{1,\sigma,\mu}(T)\right), (22)

where 2(T){\mathbb{P}}_{2}(T) is the set of all quadratic polynomials defined on T.T. This triple can be regarded as a generalization of the linear (unweighted) histopolation scheme. We employ the proposed enriched quadratic operators 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu} as building blocks for the function reconstruction method. Their quadratic nature allows them to represent curved features and ensures greater accuracy compared to the linear case.

The first step in our analysis is to show that, for any σ>0\sigma>0 and μ1\mu\geq 1, the system 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu} defines a well-posed unisolvent triple for 2(T)\mathbb{P}_{2}(T). The second step is to determine the associated basis functions. These functions are then employed to define the first two-parameter family of local function reconstruction operators.

We now summarize our main findings together with some immediate consequences. Detailed proofs will be provided at the end of this subsection.

Theorem 2.1

The triple 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu} is unisolvent for any σ>0\sigma>0 and μ1\mu\geq 1.

We next derive closed-form expressions for the basis functions associated with the enriched triple 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu}. Specifically, we determine a basis

B2,σ,μ={φi,ψi:i=1,2,3}{B}_{2,\sigma,\mu}=\left\{\varphi_{i},\psi_{i}\,:\,i=1,2,3\right\}

of the vector space 2(T)\mathbb{P}_{2}(T) satisfying the conditions

jenr(φi)\displaystyle{{\mathcal{I}}}_{j}^{\mathrm{enr}}\left(\varphi_{i}\right) =δij,\displaystyle=\delta_{ij}, (23)
jenr(φi)\displaystyle{{\mathcal{L}}}^{\mathrm{enr}}_{j}\left(\varphi_{i}\right) =0,\displaystyle=0, (24)
jenr(ψi)\displaystyle{{\mathcal{I}}}_{j}^{\mathrm{enr}}\left(\psi_{i}\right) =0,\displaystyle=0, (25)
jenr(ψi)\displaystyle{{\mathcal{L}}}^{\mathrm{enr}}_{j}\left(\psi_{i}\right) =δij,\displaystyle=\delta_{ij}, (26)

for any i,j=1,2,3i,j=1,2,3.

Theorem 2.2

The basis functions associated with the enriched triple 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu} are given by

φi\displaystyle\varphi_{i} =12λi,i=1,2,3,\displaystyle=1-2\lambda_{i},\quad i=1,2,3, (27)
ψi\displaystyle\psi_{i} =Aσ,μφi+2pσ,μ2,σ,μ2(λi2+λi+12+λi+22),i=1,2,3,\displaystyle=-A_{\sigma,\mu}\varphi_{i}+\frac{2}{\left\|{p}_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}\left(-\lambda_{i}^{2}+\lambda^{2}_{i+1}+\lambda^{2}_{i+2}\right),\quad i=1,2,3, (28)

where

pσ,μ2,σ,μ2=11pσ,μ2(t)k~σ,μ(t)𝑑t,\left\|{p}_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}=\int_{-1}^{1}{p}^{2}_{\sigma,\mu}(t)\widetilde{k}_{\sigma,\mu}(t)dt,
Aσ,μ=1+t2,σ,μpσ,μ2,σ,μ2,A_{\sigma,\mu}=\frac{1+t_{2,\sigma,\mu}}{\left\|{p}_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}, (29)

and

t2,σ,μ=11t2k~σ,μ(t)𝑑t,t_{2,\sigma,\mu}=\int_{-1}^{1}t^{2}\widetilde{k}_{\sigma,\mu}(t)dt,

denotes the second moment of k~σ,μ.\ \widetilde{k}_{\sigma,\mu}.

We now have all the necessary ingredients to define the two-parameter local reconstruction operator associated with the pdfs k~σ,μ\widetilde{k}_{\sigma,\mu}. This operator is constructed using the basis functions of Theorem 3.2 and is defined by

π1,σ,μenr:C(T)2(T)fj=13jenr(f)φj+j=13jenr(f)ψj.\begin{array}[]{rcl}{\pi}_{1,\sigma,\mu}^{{\mathrm{enr}}}:C(T)&\rightarrow&{\mathbb{P}}_{2}(T)\\ f&\mapsto&\displaystyle{\sum_{j=1}^{3}{\mathcal{I}}^{\mathrm{enr}}_{j}\left(f\right){\varphi}_{j}+\sum_{j=1}^{3}{\mathcal{L}}^{\mathrm{enr}}_{j}\left(f\right)}{\psi}_{j}.\end{array} (30)

The proofs are based on certain properties of the pdfs k~σ,μ\widetilde{k}_{\sigma,\mu} and the polynomial pσ,μp_{\sigma,\mu}. To establish Theorems 2.1 and 2.2, we first present several auxiliary lemmas and recall some concepts that will be useful in the sequel. In the discussion that follows, particular emphasis will be placed on the first three lemmas, which will be invoked repeatedly.

We begin with a lemma concerning the incomplete gamma function (3).

Lemma 1

For any ρ>0\rho>0 and s>1s>-1, the following identity holds

01tse(ρt)2𝑑t=12γmod(s+12,ρ2),\int_{0}^{1}t^{s}e^{-(\rho t)^{2}}dt=\frac{1}{2}\gamma^{\mathrm{mod}}\left(\frac{s+1}{2},\rho^{2}\right),

where γmod\gamma^{\mathrm{mod}} denotes the modified incomplete gamma function defined in (5).

Proof

By using the change of variables u=ρ2t2u=\rho^{2}t^{2}, we obtain

01tse(ρt)2𝑑t=12ρs+10ρ2us+121eu𝑑u.\int_{0}^{1}t^{s}e^{-(\rho t)^{2}}dt=\frac{1}{2\rho^{s+1}}\int_{0}^{\rho^{2}}u^{\frac{s+1}{2}-1}e^{-u}du.

The result then follows directly from the definition (5) with parameters s+12\frac{s+1}{2} and ρ2\rho^{2}.

Exploiting the symmetry of the probability density k~σ,μ\widetilde{k}_{\sigma,\mu} defined in (13), we first state a number of its fundamental properties. The next lemma shows that k~σ,μ\widetilde{k}_{\sigma,\mu} is indeed a probability density function, that is, it has unit zeroth moment.

Lemma 2

For any μ1\mu\geq 1 and σ>0\sigma>0, the function k~σ,μ\widetilde{k}_{\sigma,\mu} is a probability density function.

Proof

By symmetry and with the substitution u=tμu=t^{\mu}, we obtain

11(t2)2μ2e12(t2σ2)μ𝑑t\displaystyle\int_{-1}^{1}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}}dt =\displaystyle= 201t4μ4e(tμ2σμ)2𝑑t\displaystyle 2\int_{0}^{1}t^{4\mu-4}e^{-\left(\frac{t^{\mu}}{\sqrt{2}\sigma^{\mu}}\right)^{2}}dt (31)
=\displaystyle= 2μ01u3μ3μe(u2σμ)2𝑑u.\displaystyle\frac{2}{\mu}\int_{0}^{1}u^{\frac{3\mu-3}{\mu}}e^{-\left(\frac{u}{\sqrt{2}\sigma^{\mu}}\right)^{2}}du.

Applying Lemma 1 with

s=3μ3μ,ρ=12σμ,s=\frac{3\mu-3}{\mu},\quad\rho=\frac{1}{\sqrt{2}\sigma^{\mu}},

we can write

2μ01u3μ3μe(u2σμ)2𝑑u=1μγmod(4μ32μ,12σ2μ).\frac{2}{\mu}\int_{0}^{1}u^{\frac{3\mu-3}{\mu}}e^{-\left(\frac{u}{\sqrt{2}\sigma^{\mu}}\right)^{2}}du=\frac{1}{\mu}\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right). (32)

Then, combining (31) and (32), it follows that

11(t2)2μ2e12(t2σ2)μ𝑑t=1μγmod(4μ32μ,12σ2μ).\int_{-1}^{1}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}}dt=\frac{1}{\mu}\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right). (33)

Therefore, in order for k~σ,μ\widetilde{k}_{\sigma,\mu} to be a probability density function, the normalization constant aσ,μa_{\sigma,\mu} in (12) must be the reciprocal of the right-hand side of (33), so that

11k~σ,μ(t)𝑑t=1.\int_{-1}^{1}\widetilde{k}_{\sigma,\mu}(t)dt=1.

The next lemma allows us to derive a compact formula for the moments of k~σ,μ\widetilde{k}_{\sigma,\mu} in terms of the modified incomplete gamma functions defined in (5).

Lemma 3

Let {tm,σ,μ}m=0\left\{t_{m,\sigma,\mu}\right\}_{m=0}^{\infty} be the sequence of moments of k~σ,μ\widetilde{k}_{\sigma,\mu}. Then

tm,σ,μ=11tmk~σ,μ(t)𝑑t={0,if m=2k+1 γmod(2k+4μ32μ,12σ2μ)γmod(4μ32μ,12σ2μ),if m=2k.t_{m,\sigma,\mu}=\int_{-1}^{1}t^{m}\widetilde{k}_{\sigma,\mu}(t)dt=\begin{cases}0,&\text{if $m=2k+1$ }\\ \dfrac{\gamma^{\mathrm{mod}}\left(\dfrac{2k+4\mu-3}{2\mu},\dfrac{1}{2\sigma^{2\mu}}\right)}{\gamma^{\mathrm{mod}}\left(\dfrac{4\mu-3}{2\mu},\dfrac{1}{2\sigma^{2\mu}}\right)},&\text{if $m=2k$}.\end{cases} (34)
Proof

By symmetry, k~σ,μ\widetilde{k}_{\sigma,\mu} is even, hence for all k0k\geq 0, it follows that

t2k+1,σ,μ=0.t_{2k+1,\sigma,\mu}=0.

For the even moments, performing the change of variables u=tμu=t^{\mu} gives

t2k,σ,μ\displaystyle t_{2k,\sigma,\mu} =\displaystyle= 11t2k(t2)2μ2e12(t2σ2)μ𝑑t\displaystyle\int_{-1}^{1}t^{2k}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}}dt (35)
=\displaystyle= 201t2k+4μ4e(tμ2σμ)2𝑑t\displaystyle 2\int_{0}^{1}t^{2k+4\mu-4}e^{-\left(\frac{t^{\mu}}{\sqrt{2}\sigma^{\mu}}\right)^{2}}dt
=\displaystyle= 2μ01u2k+3μ3μe(u2σμ)2𝑑u.\displaystyle\frac{2}{\mu}\int_{0}^{1}u^{\frac{2k+3\mu-3}{\mu}}e^{-\left(\frac{u}{\sqrt{2}\sigma^{\mu}}\right)^{2}}du.

Applying Lemma 1 with

s=2k+3μ3μ,ρ=12σμ,s=\frac{2k+3\mu-3}{\mu},\quad\rho=\frac{1}{\sqrt{2}\sigma^{\mu}},

we can write

2μ01u2k+3μ3μe(u2σμ)2𝑑u=1μγmod(2k+4μ32μ,12σ2μ).\frac{2}{\mu}\int_{0}^{1}u^{\frac{2k+3\mu-3}{\mu}}e^{-\left(\frac{u}{\sqrt{2}\sigma^{\mu}}\right)^{2}}du=\frac{1}{\mu}\gamma^{\mathrm{mod}}\left(\frac{2k+4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right). (36)

Then, combining (35) and (36), it follows that

11t2k(t2)2μ2e12(t2σ2)μ𝑑t=1μγmod(2k+4μ32μ,12σ2μ).\int_{-1}^{1}t^{2k}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}}dt=\frac{1}{\mu}\gamma^{\mathrm{mod}}\left(\frac{2k+4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right). (37)

Then, by recalling the explicit expression of the pdfs k~σ,μ\widetilde{k}_{\sigma,\mu} in (13) and using (37), we obtain

t2k,σ,μ\displaystyle t_{2k,\sigma,\mu} =\displaystyle= 11t2kk~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}t^{2k}\widetilde{k}_{\sigma,\mu}(t)dt
=\displaystyle= μγmod(4μ32μ,12σ2μ)11t2k(t2)2μ2e12(t2σ2)μ𝑑t\displaystyle\frac{\mu}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}\int_{-1}^{1}t^{2k}\left(t^{2}\right)^{2\mu-2}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{\mu}}dt
=\displaystyle= γmod(2k+4μ32μ,12σ2μ)γmod(4μ32μ,12σ2μ).\displaystyle\frac{\gamma^{\mathrm{mod}}\left(\frac{2k+4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}.
Remark 3

By setting k=0k=0 in (34), one immediately recovers the normalization condition, thus confirming that k~σ,μ(t)\widetilde{k}_{\sigma,\mu}(t) is indeed a probability density function.

Remark 4

The polynomial pσ,μ{p}_{\sigma,\mu} defined in (21) can be expressed in terms of the second moment as

pσ,μ(t)=t2t2,σ,μ.p_{\sigma,\mu}(t)=t^{2}-t_{2,\sigma,\mu}. (38)

Indeed, by taking m=2m=2 in identity (34), one has

t2,σ,μ=γmod(4μ12μ,12σ2μ)γmod(4μ32μ,12σ2μ),t_{2,\sigma,\mu}=\frac{\gamma^{\mathrm{mod}}\left(\frac{4\mu-1}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{4\mu-3}{2\mu},\frac{1}{2\sigma^{2\mu}}\right)},

which directly yields (38).

The limiting moments of k~σ,μ(t)\widetilde{k}_{\sigma,\mu}(t) as σ\sigma\to\infty are easily obtained from the following remark.

Remark 5

We observe that

limσ11tmk~σ,μ(t)𝑑t=11tmk~,μ(t)𝑑t.\lim_{\sigma\to\infty}\int_{-1}^{1}t^{m}\widetilde{k}_{\sigma,\mu}(t)dt=\int_{-1}^{1}t^{m}\widetilde{k}_{\infty,\mu}(t)dt.

Indeed, from the definition (15) we have

11tmk~,μ(t)𝑑t=4μ3211tm(t2)2μ2𝑑t.\int_{-1}^{1}t^{m}\widetilde{k}_{\infty,\mu}(t)dt=\frac{4\mu-3}{2}\int_{-1}^{1}t^{m}\left(t^{2}\right)^{2\mu-2}dt.

If mm is odd, this integral vanishes by symmetry. If instead m=2km=2k, a straightforward calculation yields

11t2kk~,μ(t)𝑑t=4μ3211t2k(t2)2μ2𝑑t=4μ32k+4μ3.\int_{-1}^{1}t^{2k}\widetilde{k}_{\infty,\mu}(t)dt=\frac{4\mu-3}{2}\int_{-1}^{1}t^{2k}\left(t^{2}\right)^{2\mu-2}dt=\frac{4\mu-3}{2k+4\mu-3}.

On the other hand, using (34) together with (6), we find

t2k,,μ:=limσt2k,σ,μ=4μ32k+4μ3,t_{2k,\infty,\mu}:=\lim_{\sigma\to\infty}t_{2k,\sigma,\mu}=\frac{4\mu-3}{2k+4\mu-3},

which are exactly the moments of order 2k2k associated with the pdfs k~,μ(t)\widetilde{k}_{\infty,\mu}(t).

The following result shows that pσ,μp_{\sigma,\mu}, defined in (21), is indeed an orthogonal polynomial.

Lemma 4

The polynomial pσ,μ(t)p_{\sigma,\mu}(t) is orthogonal to all linear polynomials on [1,1][-1,1] with respect to the pdfs k~σ,μ\widetilde{k}_{\sigma,\mu}.

Proof

Since pσ,μ(t)k~σ,μ(t)p_{\sigma,\mu}(t)\widetilde{k}_{\sigma,\mu}(t) is an even function, it suffices to show that

11pσ,μ(t)k~σ,μ(t)𝑑t=0.\int_{-1}^{1}p_{\sigma,\mu}(t)\widetilde{k}_{\sigma,\mu}(t)dt=0.

Using the expression of pσ,μ(t)p_{\sigma,\mu}(t) from (38), we have

11pσ,μ(t)k~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}p_{\sigma,\mu}(t)\widetilde{k}_{\sigma,\mu}(t)dt =\displaystyle= 11(t2t2,σ,μ)k~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}\left(t^{2}-t_{2,\sigma,\mu}\right)\widetilde{k}_{\sigma,\mu}(t)dt
=\displaystyle= 11t2k~σ,μ(t)𝑑tt2,σ,μ11k~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}t^{2}\widetilde{k}_{\sigma,\mu}(t)dt-t_{2,\sigma,\mu}\int_{-1}^{1}\widetilde{k}_{\sigma,\mu}(t)dt
=\displaystyle= 0.\displaystyle 0.

In the last equality we used the facts that

11k~σ,μ(t)𝑑t=1,\int_{-1}^{1}\widetilde{k}_{\sigma,\mu}(t)dt=1,

since k~σ,μ\widetilde{k}_{\sigma,\mu} is a probability density function, and

11t2k~σ,μ(t)𝑑t=t2,σ,μ,\int_{-1}^{1}t^{2}\widetilde{k}_{\sigma,\mu}(t)dt=t_{2,\sigma,\mu},

which is precisely the second moment of k~σ,μ\widetilde{k}_{\sigma,\mu}.

We will need the average values of the barycentric coordinates on the edges of TT, weighted by the probability density k~σ,μ(t)\widetilde{k}_{\sigma,\mu}(t).

Lemma 5

For any i,j=1,2,3i,j=1,2,3, it holds

jenr(λi)=12(1δij).{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{i}\right)=\frac{1}{2}\left(1-\delta_{ij}\right). (39)
Proof

From (19), we have

jenr(λi)=11λi(1+t2𝒗j+1+1t2𝒗j+2)k~σ,μ(t)𝑑t.\mathcal{I}_{j}^{\mathrm{enr}}\left(\lambda_{i}\right)=\int_{-1}^{1}\lambda_{i}\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{k}_{\sigma,\mu}(t)dt.

If i=ji=j, then by (2) it follows immediately that

jenr(λj)=0,j=1,2,3.{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{j}\right)=0,\quad j=1,2,3.

If instead iji\neq j, then one of the values λi(𝒗j+1)\lambda_{i}(\boldsymbol{v}_{j+1}) or λi(𝒗j+2)\lambda_{i}(\boldsymbol{v}_{j+2}) equals 11, while the other equals 0. Since k~σ,μ\widetilde{k}_{\sigma,\mu} is a probability density function and barycentric coordinates are affine functions, we get

jenr(λi)\displaystyle{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{i}\right) =11(1+t2λi(𝒗j+1)+1t2λi(𝒗j+2))k~σ,μ(t)𝑑t\displaystyle=\int_{-1}^{1}\left(\frac{1+t}{2}\lambda_{i}\left(\boldsymbol{v}_{j+1}\right)+\frac{1-t}{2}\lambda_{i}\left(\boldsymbol{v}_{j+2}\right)\right)\widetilde{k}_{\sigma,\mu}(t)dt
=1211((1+t)λi(𝒗j+1)+(1t)λi(𝒗j+2))k~σ,μ(t)𝑑t\displaystyle=\frac{1}{2}\int_{-1}^{1}\left((1+t)\lambda_{i}\left(\boldsymbol{v}_{j+1}\right)+(1-t)\lambda_{i}\left(\boldsymbol{v}_{j+2}\right)\right)\widetilde{k}_{\sigma,\mu}(t)dt
=1211k~σ,μ(t)𝑑t=12,\displaystyle=\frac{1}{2}\int_{-1}^{1}\widetilde{k}_{\sigma,\mu}(t)dt=\frac{1}{2},

where in the last equality we have used the fact that k~σ,μ\widetilde{k}_{\sigma,\mu} is a probability density. Then the lemma is proved.

We also need explicit expressions for the average values of the square of the barycentric coordinates on each edge of TT, weighted by k~σ,μ(t)\widetilde{k}_{\sigma,\mu}(t).

Lemma 6

For any i,j=1,2,3i,j=1,2,3, it holds

jenr(λi2)=1+t2,σ,μ4(1δij),{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right)=\frac{1+t_{2,\sigma,\mu}}{4}\left(1-\delta_{ij}\right), (40)

where t2,σ,μt_{2,\sigma,\mu} is the second moment of k~σ,μ\widetilde{k}_{\sigma,\mu}.

Proof

From (19), we have

jenr(λi2)=11λi2(1+t2𝒗j+1+1t2𝒗j+2)k~σ,μ(t)𝑑t.{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right)=\int_{-1}^{1}\lambda_{i}^{2}\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{k}_{\sigma,\mu}(t)dt.

If i=ji=j, then by (2), it follows immediately that

jenr(λj2)=0,j=1,2,3.{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{j}^{2}\right)=0,\quad j=1,2,3.

If instead iji\neq j, we obtain

jenr(λi2)\displaystyle{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right) =11(1+t2λi(𝒗j+1)+1t2λi(𝒗j+2))2k~σ,μ(t)𝑑t\displaystyle=\int_{-1}^{1}\left(\frac{1+t}{2}\lambda_{i}\left(\boldsymbol{v}_{j+1}\right)+\frac{1-t}{2}\lambda_{i}\left(\boldsymbol{v}_{j+2}\right)\right)^{2}\widetilde{k}_{\sigma,\mu}(t)dt
=1411((1+t)λi(𝒗j+1)+(1t)λi(𝒗j+2))2k~σ,μ(t)𝑑t.\displaystyle=\frac{1}{4}\int_{-1}^{1}\left((1+t)\lambda_{i}\left(\boldsymbol{v}_{j+1}\right)+(1-t)\lambda_{i}\left(\boldsymbol{v}_{j+2}\right)\right)^{2}\widetilde{k}_{\sigma,\mu}(t)dt. (41)

Proceeding as in the proof of Lemma 5, the integral in (41) decomposes as

jenr(λi2)=1411k~σ,μ(t)𝑑t+1411t2k~σ,μ(t)𝑑t.{\mathcal{I}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right)=\frac{1}{4}\int_{-1}^{1}\widetilde{k}_{\sigma,\mu}(t)dt+\frac{1}{4}\int_{-1}^{1}t^{2}\widetilde{k}_{\sigma,\mu}(t)dt.

Since k~σ,μ\widetilde{k}_{\sigma,\mu} is a probability density, the first integral equals 11, while the second equals t2,σ,μt_{2,\sigma,\mu}. This proves (40).

The following result is an immediate consequence of Lemma 4.

Lemma 7

For any i,j=1,2,3i,j=1,2,3, the following identity holds

jenr(λi)=0.{\mathcal{L}}_{j}^{\mathrm{enr}}\left(\lambda_{i}\right)=0. (42)
Proof

The result follows directly from the definition of jenr\mathcal{L}_{j}^{\mathrm{enr}} given in (20), together with the orthogonality property of the polynomial pσ,μ(t)p_{\sigma,\mu}(t) established in Lemma 4.

Remark 6

From the definition of jenr\mathcal{L}_{j}^{\mathrm{enr}} and the orthogonality of the polynomial pσ,μp_{\sigma,\mu} established in Lemma 4, it follows that

jenr(p1)=0,j=1,2,3,\mathcal{L}_{j}^{\mathrm{enr}}\left(p_{1}\right)=0,\quad j=1,2,3,

for every linear polynomial p11(T)p_{1}\in\mathbb{P}_{1}(T).

Lemma 8

For any i,j=1,2,3i,j=1,2,3, the following identity holds

jenr(λi2)=pσ,μ2,σ,μ24(1δij).{\mathcal{L}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right)=\frac{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}{4}\left(1-\delta_{ij}\right). (43)
Proof

From equation (20) and the fact that barycentric coordinates are affine functions, we obtain

jenr(λi2)\displaystyle{\mathcal{L}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right) =11pσ,μ(t)λi2(1+t2𝒗j+1+1t2𝒗j+2)k~σ,μ(t)𝑑t\displaystyle=\int_{-1}^{1}p_{\sigma,\mu}(t)\lambda_{i}^{2}\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{k}_{\sigma,\mu}(t)dt
=11pσ,μ(t)(1+t2λi(𝒗j+1)+1t2λi(𝒗j+2))2k~σ,μ(t)𝑑t.\displaystyle=\int_{-1}^{1}p_{\sigma,\mu}(t)\left(\frac{1+t}{2}\lambda_{i}\left(\boldsymbol{v}_{j+1}\right)+\frac{1-t}{2}\lambda_{i}\left(\boldsymbol{v}_{j+2}\right)\right)^{2}\widetilde{k}_{\sigma,\mu}(t)dt.

If i=ji=j, then by (2) it follows that

jenr(λj2)=0,j=1,2,3.{\mathcal{L}}_{j}^{\mathrm{enr}}\left(\lambda_{j}^{2}\right)=0,\quad j=1,2,3.

If instead iji\neq j, then λi(𝒗j+1)=1\lambda_{i}\left(\boldsymbol{v}_{j+1}\right)=1 or λi(𝒗j+2)=1\lambda_{i}\left(\boldsymbol{v}_{j+2}\right)=1, and one of them must be zero. Using this fact, and exploiting the orthogonality property of pσ,μp_{\sigma,\mu} established in Lemma 4, together with the identity (38) expressing t2t^{2} in terms of pσ,μp_{\sigma,\mu}, we obtain

jenr(λi2)\displaystyle{\mathcal{L}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right) =1411pσ,μ(t)t2k~σ,μ(t)𝑑t\displaystyle=\frac{1}{4}\int_{-1}^{1}p_{\sigma,\mu}(t)t^{2}\widetilde{k}_{\sigma,\mu}(t)dt
=1411pσ,μ(t)(pσ,μ(t)+t2,σ,μ)k~σ,μ(t)𝑑t.\displaystyle=\frac{1}{4}\int_{-1}^{1}p_{\sigma,\mu}(t)\left(p_{\sigma,\mu}(t)+t_{2,\sigma,\mu}\right)\widetilde{k}_{\sigma,\mu}(t)dt.

Using again the orthogonality property of pσ,μp_{\sigma,\mu} the desired result follows.

We are now ready for the proof that the triple 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma,\mu} is unisolvent. To this aim, we first introduce some notations that will be used throughout the remainder of the paper. In particular, we consider the matrix

A:=[011101110].A:=\begin{bmatrix}0&1&1\\ 1&0&1\\ 1&1&0\end{bmatrix}. (44)

As det(A)=20\det(A)=2\neq 0, the matrix AA is invertible, with inverse given by

A1=12[111111111].A^{-1}=\frac{1}{2}\begin{bmatrix}-1&1&1\\ 1&-1&1\\ 1&1&-1\end{bmatrix}. (45)

Proof of Theorem 2.1. It is sufficient to prove that the only polynomial p22(T)p_{2}\in\mathbb{P}_{2}(T) satisfying

jenr(p2)\displaystyle\mathcal{I}_{j}^{\mathrm{enr}}\!\left(p_{2}\right) =0,j=1,2,3,\displaystyle=0,\quad j=1,2,3, (46)
jenr(p2)\displaystyle\mathcal{L}_{j}^{\mathrm{enr}}\!\left(p_{2}\right) =0,j=1,2,3,\displaystyle=0,\quad j=1,2,3, (47)

is the zero polynomial. To this aim, let p22(T)p_{2}\in\mathbb{P}_{2}(T), which can be expressed in barycentric form as

p2=i=13aiλi+i=13biλi2,p_{2}=\sum_{i=1}^{3}a_{i}\lambda_{i}+\sum_{i=1}^{3}b_{i}\lambda_{i}^{2}, (48)

with coefficients ai,bia_{i},b_{i}\in\mathbb{R}, i=1,2,3i=1,2,3. By Lemmas 7 and 8, condition (47) reduces to

0=jenr(p2)\displaystyle 0={\mathcal{L}}^{\mathrm{enr}}_{j}\left(p_{2}\right) =i=13bijenr(λi2)=pσ,μ2,σ,μ24i=13bi(1δij).\displaystyle=\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}b_{i}\mathcal{L}^{\mathrm{enr}}_{j}\left(\lambda^{2}_{i}\right)=\frac{\left\|p_{\sigma,\mu}\right\|^{2}_{{{2,\sigma,\mu}}}}{4}\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}b_{i}\left(1-\delta_{ij}\right).

This can equivalently be written in matrix form as

pσ,μ2,σ,μ24A𝒃=𝟎,\dfrac{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}{4}A{\boldsymbol{b}}={\boldsymbol{0}},

where AA is the matrix defined in (44), and

𝒃T=(b1,b2,b3),𝟎T=(0,0,0).{\boldsymbol{b}}^{T}=\left(b_{1},b_{2},b_{3}\right),\quad{\boldsymbol{0}}^{T}=(0,0,0).

Since AA is non-singular, we conclude that

b1=b2=b3=0.b_{1}=b_{2}=b_{3}=0.

Consequently, by (48), the polynomial p2p_{2} reduces to

p2=p1=a1λ1+a2λ2+a3λ3,a1,a2,a3.p_{2}=p_{1}=a_{1}\lambda_{1}+a_{2}\lambda_{2}+a_{3}\lambda_{3},\quad a_{1},a_{2},a_{3}\in\mathbb{R}.

Applying Lemma 5, condition (46) becomes

0\displaystyle 0 =jenr(p2)=12i=13ai(1δij).\displaystyle={\mathcal{I}}_{j}^{\mathrm{enr}}\left(p_{2}\right)=\frac{1}{2}\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}a_{i}\left(1-\delta_{ij}\right).

This can equivalently be written in matrix form as

12A𝒂=𝟎,\frac{1}{2}A{\boldsymbol{a}}={\boldsymbol{0}}, (49)

where

𝒂T=(a1,a2,a3).{\boldsymbol{a}}^{T}=\left(a_{1},a_{2},a_{3}\right).

Since AA is non-singular, system (49) admits only the trivial solution

a1=a2=a3=0,a_{1}=a_{2}=a_{3}=0,

which implies p2=0p_{2}=0 and thus completes the proof. ∎

The proof of Theorem 3.2 relies on the derivation of the explicit analytic expressions of the basis functions φi\varphi_{i} and ψi\psi_{i}. We illustrate the procedure by computing φ1\varphi_{1} and ψ1\psi_{1}, as the remaining cases follow in the same way.

Proof of Theorem 2.2.
We begin by deriving the analytic expression of the basis function φ1\varphi_{1}. According to its definition, φ1\varphi_{1} can be expressed as

φ1=i=13aiλi+i=13biλi2,\varphi_{1}=\sum_{i=1}^{3}a_{i}\lambda_{i}+\sum_{i=1}^{3}b_{i}\lambda_{i}^{2}, (50)

for some coefficients ai,bia_{i},b_{i}\in\mathbb{R}, i=1,2,3i=1,2,3. To determine these coefficients, we use condition (24) together with Lemmas 7 and 8, which yield

0=jenr(φ1)=i=13bijenr(λi2)=pσ,μ2,σ,μ24i=13bi(1δij).0=\mathcal{L}^{\mathrm{enr}}_{j}\left(\varphi_{1}\right)=\sum_{i=1}^{3}b_{i}\mathcal{L}^{\mathrm{enr}}_{j}\left(\lambda^{2}_{i}\right)=\frac{\|p_{\sigma,\mu}\|^{2}_{2,{\sigma,\mu}}}{4}\sum_{i=1}^{3}b_{i}\left(1-\delta_{ij}\right).

As in the previous proof, this system admits only the trivial solution

b1=b2=b3=0.b_{1}=b_{2}=b_{3}=0.

Substituting these values into (50), φ1\varphi_{1} reduces to

φ1=i=13aiλi.\varphi_{1}=\sum_{i=1}^{3}a_{i}\lambda_{i}.

Applying Lemma 5, condition (23) becomes

δ1j\displaystyle\delta_{1j} =jenr(φ1)=i=13aijenr(λi)=12i=13ai(1δij).\displaystyle={\mathcal{I}}^{\mathrm{enr}}_{j}\left(\varphi_{1}\right)=\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}a_{i}{\mathcal{I}}^{\mathrm{enr}}_{j}\left(\lambda_{i}\right)=\frac{1}{2}\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}a_{i}\left(1-\delta_{ij}\right).

This is equivalently written in matrix form as

(1,0,0)T=12A𝒂\left(1,0,0\right)^{T}=\frac{1}{2}A{\boldsymbol{a}}

where

𝒂T=(a1,a2,a3).{\boldsymbol{a}}^{T}=\left(a_{1},a_{2},a_{3}\right).

A straightforward computation yields

a1=1,a2=a3=1,a_{1}=-1,\quad a_{2}=a_{3}=1,

and hence

φ1=12λ1.\varphi_{1}=1-2\lambda_{1}.

It remains to compute the expression of the basis function ψ1\psi_{1}. By definition, it can be expressed as

ψ1=i=13ciλi+i=13diλi2,\psi_{1}=\sum_{i=1}^{3}c_{i}\lambda_{i}+\sum_{i=1}^{3}d_{i}\lambda_{i}^{2}, (51)

for some coefficients ci,dic_{i},d_{i}\in\mathbb{R}, i=1,2,3i=1,2,3. To determine these coefficients, we use condition (26) together with Lemmas 7 and 8, which yield

δ1j\displaystyle\delta_{1j} =jenr(ψ1)=i=13dijenr(λi2)=pσ,μ2,σ,μ24i=13di(1δij).\displaystyle={\mathcal{L}}^{\mathrm{enr}}_{j}\left(\psi_{1}\right)=\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}d_{i}{\mathcal{L}}^{\mathrm{enr}}_{j}\left(\lambda_{i}^{2}\right)=\frac{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}{4}\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}d_{i}\left(1-\delta_{ij}\right).

This is equivalently written in matrix form as

(1,0,0)T=pσ,μ2,σ,μ24A𝒅\left(1,0,0\right)^{T}=\frac{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}{4}A{\boldsymbol{d}}

where

𝒅T=(d1,d2,d3).{\boldsymbol{d}}^{T}=\left(d_{1},d_{2},d_{3}\right).

Solving this system, we have

d1=2pσ,μ2,σ,μ2,d2=d3=2pσ,μ2,σ,μ2.d_{1}=-\frac{2}{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}},\quad d_{2}=d_{3}=\frac{2}{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}.

Substituting these values into (51), ψ1\psi_{1} reduces to

ψ1=i=13ciλi+2pσ,μ2,σ,μ2(λ12+λ22+λ32).\psi_{1}=\sum_{i=1}^{3}c_{i}\lambda_{i}+\frac{2}{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}\left(-\lambda_{1}^{2}+\lambda^{2}_{2}+\lambda^{2}_{3}\right).

Applying Lemmas 5 and 6, condition (25) yields

0\displaystyle 0 =jenr(ψ1)=12i=13ci(1δij)+Aσ,μ(1δij),\displaystyle={\mathcal{I}}^{\mathrm{enr}}_{j}\left(\psi_{1}\right)=\frac{1}{2}\sum_{\begin{subarray}{c}i=1\end{subarray}}^{3}c_{i}\left(1-\delta_{ij}\right)+A_{\sigma,\mu}\left(1-\delta_{ij}\right),

where Aσ,μA_{\sigma,\mu} is defined in (29). This system admits the unique solution

c1=Aσ,μ,c2=c3=Aσ,μ.c_{1}=A_{\sigma,\mu},\quad c_{2}=c_{3}=-A_{\sigma,\mu}.

Hence ψ1\psi_{1} takes the form

ψ1=Aσ,μφ1+2pσ,μ2,σ,μ2(λ12+λ22+λ32),\psi_{1}=-A_{\sigma,\mu}\varphi_{1}+\frac{2}{\left\|p_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}\left(-\lambda_{1}^{2}+\lambda^{2}_{2}+\lambda^{2}_{3}\right),

which concludes the proof. ∎

2.4 Second family of distributions

We now present a second two-parameter family of exponential-type distributions, which generalizes the truncated normal distribution, and is given by the following bivariate function with parameters μ1\mu\geq 1 and σ>0\sigma>0

Gσ,μ(𝒙)=(2μ1)γmod(12,12σ4μ2)Hμ1(𝒙)e12(H(𝒙)σ2)2μ1,G_{\sigma,\mu}(\boldsymbol{x})=\frac{(2\mu-1)}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}H^{\mu-1}(\boldsymbol{x})e^{-\frac{1}{2}\left(\frac{H(\boldsymbol{x})}{\sigma^{2}}\right)^{2\mu-1}}, (52)

where HH is given by (8). Unlike the function of the first family, Kσ,μK_{\sigma,\mu} in (7), the function Gσ,μG_{\sigma,\mu} does not, in general, define a weight function for all values of μ\mu, because HH changes sign on TT. In fact, at the barycenter

𝒗¯:=13(𝒗1+𝒗2+𝒗3)\displaystyle\bar{\boldsymbol{v}}:=\frac{1}{3}\left(\boldsymbol{v}_{1}+\boldsymbol{v}_{2}+\boldsymbol{v}_{3}\right)

one has H(𝒗¯)=13H(\bar{\boldsymbol{v}})=-\frac{1}{3}, hence

Hμ1(𝒗¯)=(13)μ1,H^{\mu-1}(\bar{\boldsymbol{v}})=\left(-\frac{1}{3}\right)^{\mu-1},

which is negative whenever μ\mu is an integer and μ1\mu-1 is odd.

With the normalization in (52), the restriction of Gσ,μG_{\sigma,\mu} to any edge of TT yields a probability density function on [1,1][-1,1] (see Lemma 9). Indeed, using (1) and (2) we obtain

g~σ,μ(t)\displaystyle\widetilde{g}_{\sigma,\mu}(t) =Gσ,μ(1+t2𝒗j+1+1t2𝒗j+2)\displaystyle=G_{\sigma,\mu}\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)
=2μ1γmod(12,12σ4μ2)(t2)μ1e12(t2σ2)2μ1,t[1,1].\displaystyle=\frac{2\mu-1}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}\left(t^{2}\right)^{\mu-1}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{2\mu-1}},\quad t\in[-1,1]. (53)
Remark 7

When μ=1\mu=1, the edge-restricted densities g~σ,1(t)\widetilde{g}_{\sigma,1}(t) and k~σ,1(t)\widetilde{k}_{\sigma,1}(t) coincide for t[1,1]t\in[-1,1]. In particular, both reduce to the standard truncated normal distribution on [1,1][-1,1].

Refer to caption
Refer to caption
Figure 2: Normalized probability density functions g~σ,μ\widetilde{g}_{\sigma,\mu} on [1,1][-1,1]. Left: effect of the scale parameter, with fixed μ=2\mu=2 and different values of σ\sigma. Right: effect of the shape parameter, with fixed σ=0.50\sigma=0.50 and varying values of μ\mu, illustrating the increased sharpness and concentration around t=0t=0. These plots emphasize the flexibility provided by the two parameters.

Figure 2 shows that g~σ,μ\widetilde{g}_{\sigma,\mu} depends strongly on both the shape parameter μ\mu and the scale parameter σ\sigma, thereby motivating the two-parameter generalization.

To define our second two-parameter family of nonconforming histopolation operators, we introduce the following enriched linear functionals

𝒟jenr(f)\displaystyle{\mathcal{D}}_{j}^{\mathrm{enr}}(f) =11f(1+t2𝒗j+1+1t2𝒗j+2)g~σ,μ(t)𝑑t,j=1,2,3,\displaystyle=\int_{-1}^{1}f\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{g}_{\sigma,\mu}(t)dt,\quad j=1,2,3, (54)
𝒮jenr(f)\displaystyle{\mathcal{S}}_{j}^{\mathrm{enr}}\left(f\right) =11qσ,μ(t)f(1+t2𝒗j+1+1t2𝒗j+2)g~σ,μ(t)𝑑t,j=1,2,3.\displaystyle=\int_{-1}^{1}q_{\sigma,\mu}(t)f\left(\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2}\right)\widetilde{g}_{\sigma,\mu}(t)dt,\quad j=1,2,3. (55)

where

qσ,μ(t)=t2γmod(2μ+12(2μ1),12σ4μ2)γmod(12,12σ4μ2).{q}_{\sigma,\mu}(t)=t^{2}-\frac{\gamma^{\mathrm{mod}}\left(\frac{2\mu+1}{2(2\mu-1)},\frac{1}{2\sigma^{4\mu-2}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}. (56)

We then define the following triple

2,σ,μenr=(T,2(T),Σ2,σ,μenr(T)),\mathcal{H}^{\mathrm{enr}}_{2,\sigma,\mu}=\left(T,\mathbb{P}_{2}(T),\Sigma_{2,\sigma,\mu}^{\mathrm{enr}}(T)\right),

where

Σ2,σ,μenr(T)={𝒟jenr,𝒮jenr:j=1,2,3}.\Sigma^{\mathrm{enr}}_{2,\sigma,\mu}(T)=\left\{{\mathcal{D}}_{j}^{\mathrm{enr}},{{\mathcal{S}}}^{\mathrm{enr}}_{j}\,:\,j=1,2,3\right\}.
Remark 8

The distribution g~σ,μ\widetilde{g}_{\sigma,\mu} admits a beta-type (power) density as a limiting case. Indeed, using (4) with

s=12,z=12σ4μ2s=\frac{1}{2},\quad z=\frac{1}{2\sigma^{4\mu-2}}

it follows that

limz0+γmod(12,z)=limz0+γ(12,z)z=2.\lim_{z\rightarrow 0^{+}}\gamma^{\mathrm{mod}}\left(\frac{1}{2},z\right)=\lim_{z\rightarrow 0^{+}}\frac{\gamma\left(\frac{1}{2},z\right)}{\sqrt{z}}=2.

Therefore, by (53), for all t[1,1]t\in[-1,1] we obtain the pointwise limit

g~,μ(t):=limσg~σ,μ(t)=2μ12(t2)μ1.\widetilde{g}_{\infty,\mu}(t):=\lim_{\sigma\rightarrow\infty}\widetilde{g}_{\sigma,\mu}(t)=\frac{2\mu-1}{2}\left(t^{2}\right)^{\mu-1}. (57)

Hence, g~,μ\widetilde{g}_{\infty,\mu} is the symmetric beta-type (power) density on [1,1][-1,1], highlighting the limiting behaviour of the family as σ\sigma\to\infty.

Our first goal is to prove that 2,σ,μenr\mathcal{H}_{2,\sigma,\mu}^{\mathrm{enr}} is unisolvent. In a second step, we characterize the basis functions associated with this triple. Proofs are deferred to later subsections. The main results and some immediate consequences are summarized below.

Theorem 2.3

The triple 2,σ,μenr\mathcal{H}^{\mathrm{enr}}_{2,\sigma,\mu} is unisolvent for any σ>0\sigma>0 and μ1\mu\geq 1.

We next derive closed-form expressions for the basis functions associated with the triple 2,σ,μenr\mathcal{H}^{\mathrm{enr}}_{2,\sigma,\mu}. Specifically, we determine a basis

B2,σ,μ={ηi,ρi:i=1,2,3}{B}^{\prime}_{2,\sigma,\mu}=\left\{\eta_{i},\rho_{i}\,:\,i=1,2,3\right\}

of the vector space 2(T)\mathbb{P}_{2}(T) satisfying the conditions

𝒟jenr(ηi)\displaystyle{{\mathcal{D}}}_{j}^{\mathrm{enr}}\left(\eta_{i}\right) =δij,\displaystyle=\delta_{ij},
𝒮jenr(ηi)\displaystyle{{\mathcal{S}}}^{\mathrm{enr}}_{j}\left(\eta_{i}\right) =0,\displaystyle=0,
𝒟jenr(ρi)\displaystyle{\mathcal{D}}_{j}^{\mathrm{enr}}\left(\rho_{i}\right) =0,\displaystyle=0,
𝒮jenr(ρi)\displaystyle{{\mathcal{S}}}^{\mathrm{enr}}_{j}\left(\rho_{i}\right) =δij,\displaystyle=\delta_{ij},

for any i,j=1,2,3i,j=1,2,3.

Theorem 2.4

The basis functions associated with the enriched triple 2,σ,μenr\mathcal{H}^{\mathrm{enr}}_{2,\sigma,\mu} are given by

ηi\displaystyle\eta_{i} =12λi,i=1,2,3,\displaystyle=1-2\lambda_{i},\quad i=1,2,3,
ρi\displaystyle\rho_{i} =Cσ,μηi+2qσ,μ2,σ,μ2(λi2+λi+12+λi+22),i=1,2,3,\displaystyle=-C_{\sigma,\mu}\eta_{i}+\frac{2}{\left\|{q}_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}\left(-\lambda_{i}^{2}+\lambda^{2}_{i+1}+\lambda^{2}_{i+2}\right),\quad i=1,2,3,

where

qσ,μ2,σ,μ2=11qσ,μ2(t)g~σ,μ(t)𝑑t,\left\|{q}_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}=\int_{-1}^{1}{q}^{2}_{\sigma,\mu}(t)\widetilde{g}_{\sigma,\mu}(t)dt,
Cσ,μ=1+s2,σ,μqσ,μ2,σ,μ2,C_{\sigma,\mu}=\frac{1+s_{2,\sigma,\mu}}{\left\|q_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}},

and

s2,σ,μ=11t2g~σ,μ(t)𝑑t,s_{2,\sigma,\mu}=\int_{-1}^{1}t^{2}\widetilde{g}_{\sigma,\mu}(t)dt,

denotes the second moment of g~σ,μ.\ \widetilde{g}_{\sigma,\mu}.

We now have all the necessary ingredients to define the two-parameter local reconstruction operator associated with the pdfs g~σ,μ\widetilde{g}_{\sigma,\mu}. This operator is constructed using the basis functions of Theorem 2.4 and is defined by

π2,σ,μenr:C(T)2(T)fj=13𝒟jenr(f)ηj+j=13𝒮jenr(f)ρj.\begin{array}[]{rcl}{\pi}_{2,\sigma,\mu}^{{\mathrm{enr}}}:C(T)&\rightarrow&{\mathbb{P}}_{2}(T)\\ f&\mapsto&\displaystyle{\sum_{j=1}^{3}{\mathcal{D}}^{\mathrm{enr}}_{j}\left(f\right){\eta}_{j}+\sum_{j=1}^{3}{\mathcal{S}}^{\mathrm{enr}}_{j}\left(f\right)}{\rho}_{j}.\end{array}

As in the previous case, the proofs rely on specific properties of the density g~σ,μ\widetilde{g}_{\sigma,\mu} and of the polynomial qσ,μq_{\sigma,\mu}. To prove Theorems 2.3 and 2.4, we first collect a number of auxiliary lemmas and recall some standard notions needed in the sequel. Exploiting the symmetry of the edge density g~σ,μ\widetilde{g}_{\sigma,\mu} defined in (53), we begin by recording some of its basic properties. The next lemma shows that g~σ,μ\widetilde{g}_{\sigma,\mu} is a probability density on [1,1][-1,1].

Lemma 9

For any μ1\mu\geq 1 and σ>0\sigma>0, the function g~σ,μ\widetilde{g}_{\sigma,\mu} is a probability density function.

Proof

By symmetry and with the substitution u=t2μ1u=t^{2\mu-1}, we obtain

11(t2)μ1e12(t2σ2)2μ1𝑑t\displaystyle\int_{-1}^{1}\left(t^{2}\right)^{\mu-1}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{2\mu-1}}dt =\displaystyle= 201t2μ2e(t2μ12σ2μ1)2𝑑t\displaystyle 2\int_{0}^{1}t^{2\mu-2}e^{-\left(\frac{t^{2\mu-1}}{\sqrt{2}\sigma^{2\mu-1}}\right)^{2}}dt (58)
=\displaystyle= 22μ101e(u2σ2μ1)2𝑑u.\displaystyle\frac{2}{2\mu-1}\int_{0}^{1}e^{-\left(\frac{u}{\sqrt{2}\sigma^{2\mu-1}}\right)^{2}}du.

Applying Lemma 1 with

s=0,ρ=12σ2μ1,s=0,\quad\rho=\frac{1}{\sqrt{2}\sigma^{2\mu-1}},

we can write

22μ101e(uρ)2𝑑u=12μ1γmod(12,12σ4μ2).\frac{2}{2\mu-1}\int_{0}^{1}e^{-\left(\frac{u}{\rho}\right)^{2}}du=\frac{1}{2\mu-1}\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right). (59)

Then, combining (58) and (59), it follows that

11(t2)μ1e12(t2σ2)2μ1𝑑t=12μ1γmod(12,12σ4μ2).\int_{-1}^{1}\left(t^{2}\right)^{\mu-1}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{2\mu-1}}dt=\frac{1}{2\mu-1}\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right). (60)

Therefore, by the definition of g~σ,μ\widetilde{g}_{\sigma,\mu} in (53), whose normalization constant equals

2μ1γmod(12,12σ4μ2),\frac{2\mu-1}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)},

we get

11g~σ,μ(t)𝑑t=1.\int_{-1}^{1}\widetilde{g}_{\sigma,\mu}(t)dt=1.

This proves the claim.

The next lemma allows us to derive a compact formula for the moments of g~σ,μ\widetilde{g}_{\sigma,\mu} in terms of the modified incomplete gamma functions defined in (5).

Lemma 10

Let {sm,σ,μ}m=0\{s_{m,\sigma,\mu}\}_{m=0}^{\infty} be the sequence of normalized moments of g~σ,μ\widetilde{g}_{\sigma,\mu}. Then

sm,σ,μ=11tmg~σ,μ(t)𝑑t={0,if m=2k+1,γmod(2k+2μ12(2μ1),12σ4μ2)γmod(12,12σ4μ2),if m=2k.s_{m,\sigma,\mu}=\int_{-1}^{1}t^{m}\widetilde{g}_{\sigma,\mu}(t)dt=\begin{cases}0,&\text{if }m=2k+1,\\[2.84526pt] \dfrac{\gamma^{\mathrm{mod}}\left(\dfrac{2k+2\mu-1}{2(2\mu-1)},\dfrac{1}{2\sigma^{4\mu-2}}\right)}{\gamma^{\mathrm{mod}}\left(\dfrac{1}{2},\dfrac{1}{2\sigma^{4\mu-2}}\right)},&\text{if }m=2k.\end{cases} (61)
Proof

By symmetry, g~σ,μ\widetilde{g}_{\sigma,\mu} is even, hence for all k0k\geq 0, it follows that

s2k+1,σ,μ=0.s_{2k+1,\sigma,\mu}=0.

For the even moments, performing the change of variables u=t2μ1u=t^{2\mu-1} gives

s2k,σ,μ\displaystyle s_{2k,\sigma,\mu} =\displaystyle= 2μ1γmod(12,12σ4μ2)11t2k(t2)μ1e12(t2σ2)2μ1𝑑t\displaystyle\frac{2\mu-1}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}\int_{-1}^{1}t^{2k}\left(t^{2}\right)^{\mu-1}e^{-\frac{1}{2}\left(\frac{t^{2}}{\sigma^{2}}\right)^{2\mu-1}}dt
=\displaystyle= 2(2μ1)γmod(12,12σ4μ2)01t2k+2μ2e(t2μ12σ2μ1)2𝑑t\displaystyle\frac{2(2\mu-1)}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}\int_{0}^{1}t^{2k+2\mu-2}e^{-\left(\frac{t^{2\mu-1}}{\sqrt{2}\sigma^{2\mu-1}}\right)^{2}}dt
=\displaystyle= 2γmod(12,12σ4μ2)01u2k2μ1e(u2σ2μ1)2𝑑u.\displaystyle\frac{2}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}\int_{0}^{1}u^{\frac{2k}{2\mu-1}}e^{-\left(\frac{u}{\sqrt{2}\sigma^{2\mu-1}}\right)^{2}}du.

Applying Lemma 1 with

s=2k2μ1,ρ=12σ2μ1s=\frac{2k}{2\mu-1},\quad\rho=\frac{1}{\sqrt{2}\sigma^{2\mu-1}}

we can write

2γmod(12,12σ4μ2)01u2k2μ1e(u2σ2μ1)2𝑑u=γmod(2k+2μ12(2μ1),12σ4μ2)γmod(12,12σ4μ2).\frac{2}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}\int_{0}^{1}u^{\frac{2k}{2\mu-1}}e^{-\left(\frac{u}{\sqrt{2}\sigma^{2\mu-1}}\right)^{2}}du=\frac{\gamma^{\mathrm{mod}}\left(\frac{2k+2\mu-1}{2(2\mu-1)},\frac{1}{2\sigma^{4\mu-2}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}.

Therefore

s2k,σ,μ=γmod(2k+2μ12(2μ1),12σ4μ2)γmod(12,12σ4μ2).s_{2k,\sigma,\mu}=\frac{\gamma^{\mathrm{mod}}\left(\frac{2k+2\mu-1}{2(2\mu-1)},\frac{1}{2\sigma^{4\mu-2}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)}.

As an immediate application of the integral representation of the moments in Lemma 10, we obtain an alternative expression for the polynomial in (56).

Remark 9

The polynomial qσ,μ{q}_{\sigma,\mu} defined in (56) can be expressed in terms of the second moment as

qσ,μ(t)=t2s2,σ,μ.q_{\sigma,\mu}(t)=t^{2}-s_{2,\sigma,\mu}. (62)

Indeed, by taking m=2m=2 in identity (61), one has

s2,σ,μ=γmod(2μ+12(2μ1),12σ4μ2)γmod(12,12σ4μ2),s_{2,\sigma,\mu}=\frac{\gamma^{\mathrm{mod}}\left(\frac{2\mu+1}{2(2\mu-1)},\frac{1}{2\sigma^{4\mu-2}}\right)}{\gamma^{\mathrm{mod}}\left(\frac{1}{2},\frac{1}{2\sigma^{4\mu-2}}\right)},

which directly yields (62).

Remark 10

Setting m=0m=0 in (61) gives s0,σ,μ=1s_{0,\sigma,\mu}=1, confirming that g~σ,μ\widetilde{g}_{\sigma,\mu} is a probability density.

Remark 11

We observe that

limσ11tmg~σ,μ(t)𝑑t=11tmg~,μ(t)𝑑t.\lim_{\sigma\to\infty}\int_{-1}^{1}t^{m}\widetilde{g}_{\sigma,\mu}(t)dt=\int_{-1}^{1}t^{m}\widetilde{g}_{\infty,\mu}(t)dt.

Indeed, from the definition (57) we have

11tmg~,μ(t)𝑑t=2μ1211tm(t2)μ1𝑑t.\int_{-1}^{1}t^{m}\widetilde{g}_{\infty,\mu}(t)dt=\frac{2\mu-1}{2}\int_{-1}^{1}t^{m}\left(t^{2}\right)^{\mu-1}dt.

If mm is odd, this integral vanishes by symmetry. If instead m=2km=2k, a straightforward calculation yields

11t2kg~,μ(t)𝑑t=2μ1211t2k(t2)μ1𝑑t=2μ12k+2μ1.\int_{-1}^{1}t^{2k}\widetilde{g}_{\infty,\mu}(t)dt=\frac{2\mu-1}{2}\int_{-1}^{1}t^{2k}\left(t^{2}\right)^{\mu-1}dt=\frac{2\mu-1}{2k+2\mu-1}.

On the other hand, using (61) together with (6), we find

s2k,,μ:=limσs2k,σ,μ=2μ12k+2μ1,s_{2k,\infty,\mu}:=\lim_{\sigma\to\infty}s_{2k,\sigma,\mu}=\frac{2\mu-1}{2k+2\mu-1},

which are exactly the moments of order 2k2k associated with the pdfs g~,μ(t)\widetilde{g}_{\infty,\mu}(t).

The next result establishes that qσ,μq_{\sigma,\mu}, defined in (56), is the degree-two orthogonal polynomial associated with the weight g~σ,μ\widetilde{g}_{\sigma,\mu} on [1,1][-1,1].

Lemma 11

The polynomial qσ,μ(t)q_{\sigma,\mu}(t) is orthogonal to all linear polynomials on [1,1][-1,1] w.r.t. the weight function g~σ,μ\widetilde{g}_{\sigma,\mu}.

Proof

Since qσ,μ(t)g~σ,μ(t)q_{\sigma,\mu}(t)\widetilde{g}_{\sigma,\mu}(t) is an even function, it suffices to show that

11qσ,μ(t)g~σ,μ(t)𝑑t=0.\int_{-1}^{1}q_{\sigma,\mu}(t)\widetilde{g}_{\sigma,\mu}(t)dt=0. (63)

Using the expression of qσ,μq_{\sigma,\mu} from (62), we have

11qσ,μ(t)g~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}q_{\sigma,\mu}(t)\widetilde{g}_{\sigma,\mu}(t)dt =\displaystyle= 11(t2s2,σ,μ)g~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}\left(t^{2}-s_{2,\sigma,\mu}\right)\widetilde{g}_{\sigma,\mu}(t)dt (64)
=\displaystyle= 11t2g~σ,μ(t)𝑑ts2,σ,μ11g~σ,μ(t)𝑑t\displaystyle\int_{-1}^{1}t^{2}\widetilde{g}_{\sigma,\mu}(t)dt-s_{2,\sigma,\mu}\int_{-1}^{1}\widetilde{g}_{\sigma,\mu}(t)dt (65)
=\displaystyle= 0.\displaystyle 0. (66)

In the last equality we used the facts that

11g~σ,μ(t)𝑑t=1\int_{-1}^{1}\widetilde{g}_{\sigma,\mu}(t)dt=1

since g~σ,μ(t)\widetilde{g}_{\sigma,\mu}(t) is a probability density function, and

11t2g~σ,μ(t)𝑑t=s2,σ,μ\int_{-1}^{1}t^{2}\widetilde{g}_{\sigma,\mu}(t)dt=s_{2,\sigma,\mu}

which is precisely the second moment of g~σ,μ(t)\widetilde{g}_{\sigma,\mu}(t).

In what follows, we prove several crucial lemmas needed in the sequel.

Lemma 12

For any i,j=1,2,3i,j=1,2,3, it holds

𝒟jenr(λi)=12(1δij).{\mathcal{D}}_{j}^{\mathrm{enr}}\left(\lambda_{i}\right)=\frac{1}{2}\left(1-\delta_{ij}\right). (67)
Proof

The proof is similar to that of Lemma 5 and is therefore omitted.

Lemma 13

For any i,j=1,2,3i,j=1,2,3, it holds

𝒟jenr(λi2)=1+s2,σ,μ4(1δij),{\mathcal{D}}_{j}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right)=\frac{1+s_{2,\sigma,\mu}}{4}\left(1-\delta_{ij}\right), (68)

where s2,σ,μs_{2,\sigma,\mu} is the second moment of g~σ,μ\widetilde{g}_{\sigma,\mu}.

Proof

The proof is similar to that of Lemma 6 and is therefore omitted.

Lemma 14

For any i,j=1,2,3i,j=1,2,3, it holds

𝒮jenr(λi)=0.{\mathcal{S}}_{{j}}^{\mathrm{enr}}\left(\lambda_{i}\right)=0. (69)
Proof

The result follows directly from (55) and Lemma 11.

Remark 12

From the definition of 𝒮jenr\mathcal{S}_{j}^{\mathrm{enr}} and the orthogonality of the polynomial qσ,μq_{\sigma,\mu} established in Lemma 11, it follows that

𝒮jenr(p1)=0,j=1,2,3,\mathcal{S}_{j}^{\mathrm{enr}}\left(p_{1}\right)=0,\quad j=1,2,3, (70)

for every linear polynomial p11(T)p_{1}\in\mathbb{P}_{1}(T).

Lemma 15

For any i,j=1,2,3i,j=1,2,3, it holds

𝒮jenr(λi2)=qσ,μ2,σ,μ24(1δij).{\mathcal{S}}_{{j}}^{\mathrm{enr}}\left(\lambda_{i}^{2}\right)=\frac{\left\|q_{\sigma,\mu}\right\|^{2}_{2,{{\sigma,\mu}}}}{4}\left(1-\delta_{ij}\right). (71)
Proof

The proof is similar to that of Lemma 8 and is therefore omitted.

Now, we are ready to prove that the triple 2,σ,μenr\mathcal{H}_{2,\sigma,\mu}^{\mathrm{enr}} is unisolvent.

Proof of Theorem 2.3. Using Lemmas 12, 13, 14, and 15, the proof follows along the same lines as that of Theorem 2.1 and is therefore omitted. ∎

Proof of Theorem 2.4
Using Lemmas 12, 13, 14, and 15, the proof follows along the same lines as that of Theorem 2.2 and is therefore omitted.

2.5 Parameter selection by global mesh-level tuning

We discuss a complementary strategy to select the parameters of the edge densities (e.g., μ,σ\mu,\sigma for the families introduced in Section 2). Given a validation set of test functions

{f(r)}r=1R\left\{f^{(r)}\right\}_{r=1}^{R}

and a sequence of meshes

{𝒯n}n=0N,\left\{\mathcal{T}_{n}\right\}_{n=0}^{N},

we determine the optimal parameters by minimizing the global L1L^{1} error over a candidate grid 𝒫\mathcal{P} of parameter pairs:

(μ,σ)argmin(μ,σ)𝒫rnf(r)π1,σ,μenr[f(r)]L1(Ω;𝒯n),\left(\mu^{\star},\sigma^{\star}\right)\in\arg\min_{(\mu,\sigma)\in\mathcal{P}}\sum_{r}\sum_{n}\left\|f^{(r)}-{\pi}_{1,\sigma,\mu}^{{\mathrm{enr}}}\left[f^{(r)}\right]\right\|_{L^{1}(\Omega;\mathcal{T}_{n})},

where π1,σ,μenr{\pi}_{1,\sigma,\mu}^{{\mathrm{enr}}} is defined in (30). This procedure provides a robust, once-for-all choice of the parameters.

Algorithm 1 Global parameter tuning by grid search
1:Validation functions {f(r)}r=1R\left\{f^{(r)}\right\}_{r=1}^{R}; meshes {𝒯n}n=0N\{\mathcal{T}_{n}\}_{n=0}^{N}; candidate grid 𝒫\mathcal{P} of parameter pairs (μ,σ)(\mu,\sigma); reconstruction operator π1,μ,σenr\pi^{\mathrm{enr}}_{1,\mu,\sigma}
2:Optimal parameters (μ,σ)\left(\mu^{\star},\sigma^{\star}\right)
3:Emin+E_{\min}\leftarrow+\infty;  (μ,σ)undefined\left(\mu^{\star},\sigma^{\star}\right)\leftarrow\text{undefined}
4:for all (μ,σ)𝒫(\mu,\sigma)\in\mathcal{P} do
5:  E(μ,σ)0E(\mu,\sigma)\leftarrow 0
6:  for r=1r=1 to RR do
7:   for n=0n=0 to NN do
8:     uπ1,μ,σenr[f(r);𝒯n]u\leftarrow\pi^{\mathrm{enr}}_{1,\mu,\sigma}[f^{(r)};\mathcal{T}_{n}]
9:     E(μ,σ)E(μ,σ)+f(r)uL1(Ω;𝒯n)E(\mu,\sigma)\leftarrow E(\mu,\sigma)+\|f^{(r)}-u\|_{L^{1}(\Omega;\mathcal{T}_{n})}
10:   end for
11:  end for
12:  if E(μ,σ)<EminE(\mu,\sigma)<E_{\min} then
13:   EminE(μ,σ)E_{\min}\leftarrow E(\mu,\sigma);  (μ,σ)(μ,σ)\left(\mu^{\star},\sigma^{\star}\right)\leftarrow(\mu,\sigma)
14:  end if
15:end for
16:return (μ,σ)\left(\mu^{\star},\sigma^{\star}\right)

This grid–search strategy is simple to implement and guarantees the existence of a well–defined pair of optimal parameters (μ,σ)\left(\mu^{\star},\sigma^{\star}\right). Once selected, these values can be applied uniformly across all mesh elements, thus ensuring that the reconstruction procedure is completely specified and reproducible. Although the search involves multiple reconstructions on the validation set, it is performed only once as an offline preprocessing step. The modest computational cost is largely compensated by a substantial gain in robustness and accuracy in the proposed weighted histopolation framework. In practice, the method behaves like a standard hyperparameter tuning stage in machine learning, providing an automatic and principled way to calibrate the edge densities before applying the reconstruction algorithm to new data.

3 Generalization to Arbitrary Edge Probability Densities

In the previous sections, we focused on two specific two-parameter families of generalized truncated normal distributions, each giving rise to valid probability densities on the edges of a triangular element. A key observation is that, regardless of the chosen family, the resulting edge-restricted densities share the same structural properties: they are normalized probability densities, symmetric with respect to the edge parameter, and admit finite moments of all orders. These properties alone are sufficient to establish unisolvence of the enriched triple and to construct explicit quadratic reconstruction operators.

This motivates us to consider a more general setting. Let T2T\subset\mathbb{R}^{2} be a nondegenerate triangle with edges sj=[𝒗j+1,𝒗j+2]s_{j}=\left[\boldsymbol{v}_{j+1},\boldsymbol{v}_{j+2}\right], and denote by

γj(t)=1+t2𝒗j+1+1t2𝒗j+2,t[1,1],\gamma_{j}(t)=\frac{1+t}{2}\boldsymbol{v}_{j+1}+\frac{1-t}{2}\boldsymbol{v}_{j+2},\quad t\in[-1,1],

the affine parametrization of the edge sjs_{j}. Let ωL1([1,1])\omega\in L^{1}([-1,1]) be a general pdf on [1,1][-1,1]. We assume further that ω\omega admits finite moments

μm,ω=11tmω(t)𝑑t,m.\mu_{m,\omega}=\int_{-1}^{1}t^{m}\omega(t)dt,\quad m\in\mathbb{N}. (72)

3.1 Edge functionals and orthogonal polynomials

Given such a density ω\omega, we define enriched edge functionals in analogy with the previous constructions by

jω(f)\displaystyle\mathcal{I}^{\omega}_{j}(f) =11f(γj(t))ω(t)𝑑t,\displaystyle=\int_{-1}^{1}f(\gamma_{j}(t))\omega(t)dt, (73)
jω(f)\displaystyle\mathcal{L}^{\omega}_{j}(f) =11q(t)f(γj(t))ω(t)𝑑t,\displaystyle=\int_{-1}^{1}q(t)f(\gamma_{j}(t))\omega(t)dt, (74)

where qr([1,1])q\in\mathbb{P}_{r}([-1,1]), r2r\geq 2, is a polynomial satisfying the orthogonality conditions

11q(t)ω(t)𝑑t=0,11tq(t)ω(t)𝑑t=0,\int_{-1}^{1}q(t)\omega(t)dt=0,\quad\int_{-1}^{1}tq(t)\omega(t)dt=0, (75)

together with the non-degeneracy condition on the second moment

μ2,ω:=11t2q(t)ω(t)𝑑t0.\mu_{2,\omega}:=\int_{-1}^{1}t^{2}q(t)\omega(t)dt\neq 0. (76)

These assumptions ensure that the qq-weighted edge functionals (74) vanish on affine polynomials while detecting quadratic contributions.

3.2 Unisolvency and reconstruction

We then define the enriched triple

ω=(T,2(T),Σω(T)),\mathcal{H}^{\omega}=\left(T,\mathbb{P}_{2}(T),\Sigma^{\omega}(T)\right),

where

Σω(T)={jω,jω:j=1,2,3}.\Sigma^{\omega}(T)=\left\{\mathcal{I}^{\omega}_{j},\mathcal{L}^{\omega}_{j}\,:\,j=1,2,3\right\}.

The following result shows that the structural properties observed for the specific families extend to any choice of probability density ω\omega.

Theorem 3.1

The triple ω\mathcal{H}^{\omega} is unisolvent.

Proof(Sketch of proof)

The argument follows the same lines as in the proofs of Theorems 2.3 and 2.19. Indeed, let p22(T)p_{2}\in\mathbb{P}_{2}(T) such that

jω(p2)=jω(p2)=0,j=1,2,3.\mathcal{I}_{j}^{\omega}\left(p_{2}\right)=\mathcal{L}_{j}^{\omega}\left(p_{2}\right)=0,\quad j=1,2,3.

Along each edge sjs_{j}, the trace of p2p_{2} is quadratic. By construction, jω\mathcal{L}^{\omega}_{j} annihilates constant and linear terms, while detecting the quadratic coefficient. Hence, vanishing of jω(p2)\mathcal{L}^{\omega}_{j}\left(p_{2}\right) implies that all edge traces of p2p_{2} are affine. This forces p21(T)p_{2}\in\mathbb{P}_{1}(T). The remaining conditions jω(p2)=0\mathcal{I}^{\omega}_{j}\left(p_{2}\right)=0 then yield a homogeneous system for the linear coefficients, which admits only the trivial solution. This completes the proof.

This generalization shows that the enriched quadratic reconstruction framework is not restricted to the two-parameter families considered earlier, but in fact applies to any choice of probability density on the edges. The specific distributions k~σ,μ\widetilde{k}_{\sigma,\mu} and g~σ,μ\widetilde{g}_{\sigma,\mu} therefore appear as special cases within a broad and flexible class, highlighting the robustness and universality of the proposed approach.

Theorem 3.2

If the probability density function ω\omega is even, then the basis functions associated with the enriched triple ω\mathcal{H}^{\omega} take the form

φω,i\displaystyle\varphi_{\omega,i} =12λi,i=1,2,3,\displaystyle=1-2\lambda_{i},\quad i=1,2,3,
ψω,i\displaystyle\psi_{\omega,i} =Aωφω,i+2q2,ω2(λi2+λi+12+λi+22),i=1,2,3,\displaystyle=-A_{\omega}\varphi_{\omega,i}+\frac{2}{\left\|{q}\right\|^{2}_{2,{\omega}}}\left(-\lambda_{i}^{2}+\lambda^{2}_{i+1}+\lambda^{2}_{i+2}\right),\quad i=1,2,3,

where

q2,ω2=11q2(t)ω(t)𝑑t,\left\|{q}\right\|^{2}_{2,{\omega}}=\int_{-1}^{1}q^{2}(t)\omega(t)dt,
Aω=1+μ2,ωq2,ω2.A_{\omega}=\frac{1+\mu_{2,\omega}}{\left\|{q}\right\|^{2}_{2,{\omega}}}.

3.3 Practical constructions of qq

When qq is chosen to be a quadratic polynomial, it is possible to construct it explicitly from the moments μm,ω\mu_{m,\omega}. We outline below three practical approaches.

  1. (A)

    Canonical quadratic form. Seek q(t)=t2abtq(t)=t^{2}-a-bt subject to the orthogonality conditions (75). Using the notation (72), the orthogonality conditions yield the linear system

    {μ2,ωabμ1,ω=0,μ3,ωaμ1,ωbμ2,ω=0,b=μ1,ωμ2,ωμ3,ωμ1,ω2μ2,ω,a=μ2,ωbμ1,ω,\begin{cases}\mu_{2,\omega}-a-b\mu_{1,\omega}=0,\\ \mu_{3,\omega}-a\mu_{1,\omega}-b\mu_{2,\omega}=0,\end{cases}\quad\Longrightarrow\quad b=\frac{\mu_{1,\omega}\mu_{2,\omega}-\mu_{3,\omega}}{\mu_{1,\omega}^{2}-\mu_{2,\omega}},\quad a=\mu_{2,\omega}-b\mu_{1,\omega},

    provided μ2,ωμ1,ω2\mu_{2,\omega}\neq\mu_{1,\omega}^{2} (nonzero variance). Optionally normalize qq by

    q^=qν\widehat{q}=\frac{q}{\nu}

    where

    ν=μ4,ωaμ2,ωbμ3,ω.\nu=\mu_{4,\omega}-a\mu_{2,\omega}-b\mu_{3,\omega}.
  2. (B)

    Orthogonal polynomial. Construct the family of orthogonal polynomials {πk}k0\left\{\pi_{k}\right\}_{k\geq 0} with respect to the inner product

    f,gω=11f(t)g(t)ω(t)𝑑t.\left\langle f,g\right\rangle_{\omega}=\int_{-1}^{1}f(t)g(t)\omega(t)dt.

    Taking q=π2q=\pi_{2} (or its normalized version) ensures that conditions (75)–(76) are satisfied by construction milovanovic1997orthogonal; mastroianni2008interpolation.

  3. (C)

    Higher degree. Any polynomial qq of degree 2\geq 2 fulfilling (75)–(76) is admissible. The quadratic choice is however preferred, since it requires only low-order moments and ensures stability.

Remark 13

If the chosen density ω\omega is symmetric with respect to the origin, then all odd moments vanish, i.e.,

μ2k+1,ω=0,k.\mu_{2k+1,\omega}=0,\quad k\in\mathbb{N}.

In this case, the expressions in (A) simplify considerably: one obtains b=0b=0 and hence

q(t)=t2μ2,ω.q(t)=t^{2}-\mu_{2,\omega}.

This situation arises in particular for both the first and the second family of densities introduced in Section 2, which are symmetric and therefore lead to especially simple quadratic forms.

3.4 Numerical procedure

Before presenting the numerical results, we summarize the computational workflow adopted in our experiments. The procedure consists of two phases:

  • Offline parameter tuning: optimal density parameters (μ,σ)\left(\mu^{\star},\sigma^{\star}\right) are selected by Algorithm 1, applied to a small validation set of functions and a sequence of meshes.

  • Reconstruction and error evaluation: once the optimal parameters are fixed, they are used uniformly across all tests to compare the classical histopolation scheme 𝒞\mathcal{CH} with the enriched scheme 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}}.

The pseudocode of the complete workflow is reported in Algorithm 2.

Algorithm 2 Numerical testing workflow
1:Test functions {fr}r=1R\left\{f_{r}\right\}_{r=1}^{R}; mesh sequence {𝒯n}n=0N\left\{\mathcal{T}_{n}\right\}_{n=0}^{N}; parameter grid 𝒫\mathcal{P}
2:L1L^{1} errors for 𝒞\mathcal{CH} and 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}}
3:(μ,σ)\left(\mu^{\star},\sigma^{\star}\right)\leftarrow parameter tuning via Algorithm 1
4:for each test function fjf_{j} do
5:  for each mesh 𝒯n\mathcal{T}_{n} do
6:   uCHπCH[fj;𝒯n]u^{\mathrm{CH}}\leftarrow\pi^{\mathrm{CH}}\left[f_{j};\mathcal{T}_{n}\right]
7:   uenrπ1,σ,μenr[fj;𝒯n]u^{\mathrm{enr}}\leftarrow\pi_{1,\sigma,\mu}^{\mathrm{enr}}\left[f_{j};\mathcal{T}_{n}\right]
8:   Compute Ej,nCH=fjuCHL1(Ω;𝒯n)E^{\mathrm{CH}}_{j,n}=\left\|f_{j}-u^{\mathrm{CH}}\right\|_{L^{1}(\Omega;\mathcal{T}_{n})}
9:   Compute Ej,nenr=fjuenrL1(Ω;𝒯n)E^{\mathrm{enr}}_{j,n}=\left\|f_{j}-u^{\mathrm{enr}}\right\|_{L^{1}(\Omega;\mathcal{T}_{n})}
10:  end for
11:end for
12:return Error arrays {Ej,nCH,Ej,nenr}\left\{E^{\mathrm{CH}}_{j,n},E^{\mathrm{enr}}_{j,n}\right\}

This explicit workflow highlights how the parameter tuning step integrates seamlessly into the testing phase: after a one-time offline calibration, the enriched scheme is applied to new test functions and meshes in a fully specified and reproducible way.

4 Numerical tests

In this section, we present a series of numerical experiments to assess the performance and accuracy of the proposed enriched histopolation methods. All tests are carried out on the square domain Ω=[1,1]2\Omega=[-1,1]^{2}, using a collection of benchmark functions that exhibit a variety of analytic behaviors, including smooth, oscillatory, and localized features. The test functions are defined as follows

f1(x,y)\displaystyle f_{1}(x,y) =\displaystyle= x2+y2,f2(x,y)=e4(x2+y2)sin(π(x+y)),\displaystyle\sqrt{x^{2}+y^{2}},\ \ f_{2}(x,y)=e^{-4\left(x^{2}+y^{2}\right)}\sin\left(\pi(x+y)\right),
f3(x,y)\displaystyle f_{3}(x,y) =\displaystyle= sin(2πx)sin(2πy),f4(x,y)=sin(4π(x+y)),f5(x,y)=125(x2+y2)+1,\displaystyle\sin(2\pi x)\sin(2\pi y),\ \,f_{4}(x,y)=\sin\left(4\pi(x+y)\right),\ \,f_{5}(x,y)=\frac{1}{25(x^{2}+y^{2})+1},
f6(x,y)\displaystyle f_{6}(x,y) =\displaystyle= 0.75exp((9(x+1)/22)24(9(y+1)/22)24)\displaystyle 0.75\exp\biggl(-\frac{(9(x+1)/2-2)^{2}}{4}-\frac{(9(y+1)/2-2)^{2}}{4}\biggr)
+0.75exp((9(x+1)/2+1)249(9(y+1)/2+1)10)\displaystyle+0.75\exp\biggl(-\frac{(9(x+1)/2+1)^{2}}{49}-\frac{(9(y+1)/2+1)}{10}\biggr)
+0.5exp((9(x+1)/27)24(9(y+1)/23)24)\displaystyle+0.5\exp\biggl(-\frac{(9(x+1)/2-7)^{2}}{4}-\frac{(9(y+1)/2-3)^{2}}{4}\biggr)
0.2exp((9(x+1)/24)2(9(y+1)/27)2).\displaystyle-0.2\exp\biggl(-(9(x+1)/2-4)^{2}-(9(y+1)/2-7)^{2}\biggr).

The function f6f_{6} is the well-known Franke function, widely used as a benchmark for approximation methods franke1982scattered. For the spatial discretization, we consider a sequence of regular Friedrichs–Keller triangulations Knabner, denoted by

𝒯n={Ti:i=1,,2(n+1)2},\mathcal{T}_{n}=\left\{T_{i}:i=1,\dots,2(n+1)^{2}\right\},

where each 𝒯n\mathcal{T}_{n} consists of 2(n+1)22(n+1)^{2} triangles; see Fig. 3.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Successive regular Friedrichs–Keller triangulations 𝒯n\mathcal{T}_{n}, with n=20n=20 (top left), n=30n=30 (top right), n=40n=40 (bottom left), and n=50n=50 (bottom right).

To evaluate the quality of the reconstruction, we compute the error in the L1L^{1} norm produced by the classical linear histopolation method 𝒞\mathcal{CH} defined in (17) and compare it with the error obtained using the enriched histopolation method 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}} introduced in (22), with the optimal parameters μ=2\mu^{\star}=2 and σ=1\sigma^{\star}=1. These parameters have been determined by means of the parameter tuning Algorithm 1. All experiments have been carried out according to Algorithm 2, implemented in Matlab. The integrals needed for assembling the reconstruction operators and for evaluating the L1L^{1} errors were computed on each triangular element using high-order quadrature rules, with a sufficiently dense set of nodes to ensure accurate numerical integration and to rule out any artifacts due to quadrature errors. The L1L^{1} norm is chosen as it handles discontinuities effectively without requiring limiting procedures (cf. dell2025truncated). The results of this comparison are reported in Figures 45 and 6. A clear improvement of the enriched method over the standard scheme is observed across all test functions and mesh refinements. In particular, 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}} achieves significantly smaller L1L^{1} errors, demonstrating its superior capability in reproducing local oscillations, sharp gradients, and singular features. Moreover, the enriched approach demonstrates an accelerated error decay under mesh refinement, thereby confirming that the additional degrees of freedom provided by the probability density weights yield a genuine enhancement in approximation power.

Refer to caption
Refer to caption
Figure 4: Semi-log plot of the L1L^{1} approximation error for f1f_{1} (left) and f2f_{2} (right) obtained with the classical histopolation method 𝒞\mathcal{CH} (blue) and the weighted quadratic enriched method 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}} (red), as the number of triangles in the Friedrichs–Keller triangulations increases.
Refer to caption
Refer to caption
Figure 5: Semi-log plot of the L1L^{1} approximation error for f3f_{3} (left) and f4f_{4} (right) obtained with the classical histopolation method 𝒞\mathcal{CH} (blue) and the weighted quadratic enriched method 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}} (red), as the number of triangles in the Friedrichs–Keller triangulations increases.
Refer to caption
Refer to caption
Figure 6: Semi-log plot of the L1L^{1} approximation error for f5f_{5} (left) and f6f_{6} (right) obtained with the classical histopolation method 𝒞\mathcal{CH} (blue) and the weighted quadratic enriched method 1,σ,μenr\mathcal{H}^{\mathrm{enr}}_{1,\sigma^{\star},\mu^{\star}} (red), as the number of triangles in the Friedrichs–Keller triangulations increases.

5 Conclusions and Future works

In this work we have introduced two new two-parameter families of generalized truncated normal distributions and exploited them to design enriched local histopolation schemes based on unisolvent triples. These distributions serve as flexible edge weights and provide additional degrees of freedom, which translate into improved accuracy and adaptability of the reconstruction operators. The resulting quadratic operators preserve the simplicity of the classical linear histopolation method while significantly enhancing their approximation capabilities, especially for functions with oscillations, steep gradients, or localized singularities.

From a theoretical point of view, we have established unisolvency and derived explicit closed-form basis functions, thus ensuring that the proposed schemes are mathematically well-posed and computationally viable. We have also proposed an algorithm for the optimal selection of the distribution parameters, which further increases the robustness and adaptivity of the approach. Numerical experiments have confirmed the theoretical findings, showing systematic improvements over the linear nonconforming histopolation approach. Moreover, as shown in Section 3, the same reasoning extends to any bivariate weight function whose restriction to the edges of the triangle defines a valid probability density. This highlights the generality and robustness of the proposed framework, which is not limited to the two specific families presented in this paper.

Several directions for future research naturally arise from this study. First, extending the framework to three-dimensional meshes would broaden its applicability to volumetric data reconstruction and tomographic imaging. Second, adaptive strategies for selecting the distribution parameters could be developed, so as to optimize accuracy depending on local features of the target function. Third, connections with other classes of orthogonal polynomials and probability densities could be investigated, possibly leading to new families of enriched operators. Finally, applications to real-world problems in imaging, computer vision, and numerical approximation remain a promising avenue for further exploration.

Acknowledgments

This research has been achieved as part of RITA “Research ITalian network on Approximation” and as part of the UMI group “Teoria dell’Approssimazione e Applicazioni”. The research was supported by GNCS–INdAM 2025 project “Polinomi, Splines e Funzioni Kernel: dall’Approssimazione Numerica al Software Open-Source”. The work of F. Nudo is funded from the European Union – NextGenerationEU under the Italian National Recovery and Resilience Plan (PNRR), Mission 4, Component 2, Investment 1.2 “Finanziamento di progetti presentati da giovani ricercatori”, pursuant to MUR Decree No. 47/2025. The research was supported by by the grant “Bando Professori visitatori 2022” which has allowed the visit of Prof. Allal Guessab to the Department of Mathematics and Computer Science of the University of Calabria in the spring 2022. The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through Large Research Project under grant number RGP2/305/46.

Conflict of interest

Not Applicable.