[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (18)

Search Parameters:
Keywords = copula trees

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 4258 KiB  
Article
Covering Blue Voices: African American English and Authenticity in Blues Covers
by Romeo De Timmerman and Stef Slembrouck
Languages 2024, 9(7), 229; https://doi.org/10.3390/languages9070229 - 25 Jun 2024
Viewed by 629
Abstract
Many musicologists and researchers of popular music have recently stressed the omnipresence of covers in today’s music industry. In the sociolinguistics of music, however, studio-recorded covers and their potential differences from ‘original’ compositions have certainly been acknowledged in passing, but very few sociolinguists [...] Read more.
Many musicologists and researchers of popular music have recently stressed the omnipresence of covers in today’s music industry. In the sociolinguistics of music, however, studio-recorded covers and their potential differences from ‘original’ compositions have certainly been acknowledged in passing, but very few sociolinguists concerned with the study of song seem to have systematically explored how language use may differ in such re-imagined musical outputs. This article reports on a study which examines the language use of 45 blues artists from three distinct time periods (viz., 1960s, 1980s, and 2010s) and three specific social groups (viz., African American; non-African American, US-based; and non-African American, non-US based) distributed over 270 studio-recorded original and cover performances. Through gradient boosting decision tree classification, it aims to analyze the artists’ use of eight phonological and lexico-grammatical features that are traditionally associated with African American English (viz., /aɪ/ monophthongization, post-consonantal word-final /t/ deletion, post-consonantal word-final /d/ deletion, alveolar nasal /n/ in <ing> ultimas, post-vocalic word-final /r/ deletion, copula deletion, third-person singular <s> deletion, and not-contraction). Our analysis finds song type (i.e., the distinction between covers and originals) to have no meaningful impact on artists’ use of the examined features of African American English. Instead, our analysis reveals how performers seem to rely on these features to a great extent and do so markedly consistently, regardless of factors such as time period, socio-cultural background, or song type. This paper hence builds on our previous work on the language use of blues performers by further teasing out the complex indexical and iconic relationships between features of African American English, authenticity, and the blues genre in its various manifestations of time, place, and performance types. Full article
(This article belongs to the Special Issue Interface between Sociolinguistics and Music)
Show Figures

Figure 1

Figure 1
<p>Mean AAE realizations by artist and by group. Dots represent individual artist means; dashed vertical lines show group means. Abbreviations are ‘AA’ for African American; ‘nonAA_US’ for non-African American, US-based; and ‘nonAA_nonUS’ for non-African American, non-US-based.</p>
Full article ">Figure 2
<p>Point plots of mean AAE realizations by variable, group, and song type. The color blue is used for cover songs and orange for originals. Horizontal lines show 95% confidence intervals. Abbreviations are ‘AA’ for African American; ‘nonAA_US’ for non-African American, US-based; and ‘nonAA_nonUS’ for non-African American, non-US-based.</p>
Full article ">Figure 3
<p>Point plots of mean AAE realizations by artist, group, and song type. The color blue is used for cover songs, orange for originals. Horizontal lines show 95% confidence intervals. Abbreviations are ‘AA’ for African American; ‘nonAA_US’ for non-African American, US-based; and ‘nonAA_nonUS’ for non-African American, non-US-based.</p>
Full article ">Figure 4
<p>Confusion matrix for model evaluation. Displayed counts are based on model predictions on the test set (15% of data, not seen by model during training).</p>
Full article ">Figure 5
<p>Mean absolute Shapley values for all variables. All values are in log-odds space.</p>
Full article ">Figure 6
<p>Shapley values for datapoint id = 141. The mean Shapley value, E[F(X)], represents the average prediction, while f(X) indicates the prediction for this specific datapoint. The sum of the Shapley values of the individual features equals f(X). All values are in log-odds space.</p>
Full article ">Figure 7
<p>Shapley values for datapoint id = 262. The mean Shapley value, E[F(X)], represents the average prediction, while f(X) indicates the prediction for this specific datapoint. The sum of the Shapley values of the individual features equals f(X). All values are in log-odds space.</p>
Full article ">Figure 8
<p>Boxplots of Shapley values for each AAE feature. Orange lines show median value. Dots indicate outliers. All values are in log-odds space.</p>
Full article ">
13 pages, 6704 KiB  
Article
A Copula Approach for Predicting Tree Sap Flow Based on Vapor Pressure Deficit
by Ying Ouyang and Changyou Sun
Forests 2024, 15(4), 695; https://doi.org/10.3390/f15040695 - 13 Apr 2024
Viewed by 960
Abstract
While using sap-flow sensor measurements is a well-established technique for quantifying leaf water transpiration in tree species, installing and maintaining a large number of sensors and data loggers in large-scale plantations to obtain accurate measurements is both costly and time-consuming. We developed a [...] Read more.
While using sap-flow sensor measurements is a well-established technique for quantifying leaf water transpiration in tree species, installing and maintaining a large number of sensors and data loggers in large-scale plantations to obtain accurate measurements is both costly and time-consuming. We developed a copula-based approach to predict sap flows based on readily available vapor pressure deficits (VPDs) and found that the Normal copula was the best among five commonly used copulas. The Normal-copula approach was validated using our field-measured eastern cottonwood (Populus deltoides (Bartr. ex Marsh.)) sap flow data, yielding solid statistical measures, including Mann–Kendall’s τ = 0.59, R2 = 0.81, and p-value < 0.01. The approach was applied to predict sap flows of eastern cottonwood during the growing period from 1 March to 31 October 2015 as well as the 5-year growing period from 2019 to 2023. It successfully replicated the characteristic diurnal sap flow pattern, with rates increasing during the day and decreasing at night, as well as the typical seasonal pattern, with rates rising from winter to summer and decreasing from summer to next winter. Our study suggests that the copula-based approach is a reliable tool for estimating sap flows based on VPD data. Full article
(This article belongs to the Special Issue Hydrological Modelling of Forested Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Location of the Hollandale biomass study site in Mississippi, USA.</p>
Full article ">Figure 2
<p>Relationship between measured sap flow and VPD (<b>a</b>) and diurnal patterns of the measured sap flow and VPD over a week in spring 2016 (<b>b</b>).</p>
Full article ">Figure 3
<p>Percentages of maximum peak values for sap flow and VPD occurred in 24 h (<b>a</b>) and their adjustment (<b>b</b>).</p>
Full article ">Figure 4
<p>Histogram plots of VPD (<b>a</b>) and sap flow (<b>b</b>) data, showing a Gamma type of distribution.</p>
Full article ">Figure 5
<p>Flow chart used to perform copula analysis.</p>
Full article ">Figure 6
<p>Comparison of the measured and copula-predicted sap flows and VPDs. The Mann–Kendall’s <span class="html-italic">τ</span> is 0.59 between the predicted and measured sap flows at <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Relationship between the Normal copula predicted sap flow and VPD.</p>
Full article ">Figure 8
<p>Comparison of the predicted sap flows using the copula regression Equation (4) with the measured sap flows.</p>
Full article ">Figure 9
<p>Copula predicted hourly sap flows over a growing season (<b>a</b>) and a 2-week period during summer (<b>b</b>).</p>
Full article ">Figure 10
<p>Copula predicted daily (<b>a</b>) and seasonal (<b>b</b>) sap flows over a 5-year period from 2019 to 2023.</p>
Full article ">
28 pages, 5733 KiB  
Article
A Framework for Analyzing Individual-Tree and Whole-Stand Growth by Fusing Multilevel Data: Stochastic Differential Equation and Copula Network
by Petras Rupšys, Gintautas Mozgeris, Edmundas Petrauskas and Ričardas Krikštolaitis
Forests 2023, 14(10), 2037; https://doi.org/10.3390/f14102037 - 11 Oct 2023
Viewed by 925
Abstract
In forestry, growth functions form the basis of research and are widely used for the mathematical modeling of stand variables, e.g., tree or stand basal area, stand height, stand volume, site index, and many more. In this study, to estimate five-dimensional dependencies between [...] Read more.
In forestry, growth functions form the basis of research and are widely used for the mathematical modeling of stand variables, e.g., tree or stand basal area, stand height, stand volume, site index, and many more. In this study, to estimate five-dimensional dependencies between tree diameter at breast height, potentially available area, height, crown area and crown base height, we used a normal copula approach whereby the growths of individual variables are described using a stochastic differential equation with mixed-effect parameters. The normal copula combines the marginal distributions of tree diameter at breast height, potentially available area, height, crown area, and crown base height into a joint multivariate probability distribution. Copula models have the advantage of being able to use collected longitudinal, multivariate, and discrete data for which the number of measurements of individual variables does not match. This study introduced a normalized multivariate interaction information measure based on differential entropy to assess the causality between tree size variables. In order to accurately and quantitatively assess the stochastic processes of the tree size variables’ growth and to provide a scientific basis for the formalization of models, an analysis method of the synergetic theory of information entropy has been proposed. Theoretical findings are illustrated using an uneven-aged, mixed-species empirical dataset of permanent experimental plots in Lithuania. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Scheme of the algorithm of the growth models (SDE—stochastic differential equation, pdf—probability density function, AMLP—approximated maximum likelihood procedure, and MLP—maximum likelihood procedure).</p>
Full article ">Figure 2
<p>Classification of obtained measurements.</p>
Full article ">Figure 3
<p>Probability (in percent) of the event that the tree size variable will take a value from a certain interval for pine trees (red color), spruce trees (green color), and birch trees (black color): (<b>a</b>) trees in the stand have a diameter of at least 20 cm; (<b>b</b>) trees in the stand have a potentially occupied area no less than 10 m<sup>2</sup>; (<b>c</b>) trees in the stand have a height of at least 20 m; (<b>d</b>) trees in the stand have a crown base height of no less than 10 cm; (<b>e</b>) trees in the stand have a crown area of at least 20 m<sup>2</sup>.</p>
Full article ">Figure 4
<p>Mean (solid line), median (dotted line), mode (dashed line), lower quantile (dot–dashed line), and upper quantile (dot–dashed line) dynamics for pine trees (red color), spruce trees (green color), and birch trees (black color): (<b>a</b>) growth process of tree diameter; (<b>b</b>) growth process of tree occupied area; (<b>c</b>) growth process of tree height; (<b>d</b>) growth process of tree crown base height; (<b>e</b>) growth process of tree crown area; observed values—circles.</p>
Full article ">Figure 4 Cont.
<p>Mean (solid line), median (dotted line), mode (dashed line), lower quantile (dot–dashed line), and upper quantile (dot–dashed line) dynamics for pine trees (red color), spruce trees (green color), and birch trees (black color): (<b>a</b>) growth process of tree diameter; (<b>b</b>) growth process of tree occupied area; (<b>c</b>) growth process of tree height; (<b>d</b>) growth process of tree crown base height; (<b>e</b>) growth process of tree crown area; observed values—circles.</p>
Full article ">Figure 5
<p>Dynamic of live crown ratio for three randomly selected plots (the first plot—red color, species composition—P 42.5, E 51.1, B 6.4; the second plot—black color, species composition—P 49.1, E 28.1, B 22.8; the third plot—green color, species composition—P 88.9, E 0.0, B 11.1): (<b>a</b>) pine trees; (<b>b</b>) spruce trees; (<b>c</b>) birch trees; observed values—circles.</p>
Full article ">
19 pages, 920 KiB  
Article
A Statistical Dependence Framework Based on a Multivariate Normal Copula Function and Stochastic Differential Equations for Multivariate Data in Forestry
by Ričardas Krikštolaitis, Gintautas Mozgeris, Edmundas Petrauskas and Petras Rupšys
Axioms 2023, 12(5), 457; https://doi.org/10.3390/axioms12050457 - 8 May 2023
Cited by 2 | Viewed by 1226
Abstract
Stochastic differential equations and Copula theories are important topics that have many advantages for applications in almost every discipline. Many studies in forestry collect longitudinal, multi-dimensional, and discrete data for which the amount of measurement of individual variables does not match. For example, [...] Read more.
Stochastic differential equations and Copula theories are important topics that have many advantages for applications in almost every discipline. Many studies in forestry collect longitudinal, multi-dimensional, and discrete data for which the amount of measurement of individual variables does not match. For example, during sampling experiments, the diameters of all trees, the heights of approximately 10% of the trees, and the tree crown base height and crown width for a significantly smaller number of trees are measured. In this study, for estimating five-dimensional dependencies, we used a normal copula approach, where the dynamics of individual tree variables (diameter, potentially available area, height, crown base height, and crown width) are described by a stochastic differential equation with mixed-effect parameters. The approximate maximum likelihood method was used to obtain parameter estimates of the presented stochastic differential equations, and the normal copula dependence parameters were estimated using the pseudo-maximum likelihood method. This study introduced the normalized multi-dimensional interaction information index based on differential entropy to capture dependencies between state variables. Using conditional copula-type probability density functions, the exact form equations defining the links among the diameter, potentially available area, height, crown base height, and crown width were derived. All results were implemented in the symbolic algebra system MAPLE. Full article
(This article belongs to the Section Mathematical Analysis)
Show Figures

Figure 1

Figure 1
<p>Conditional and marginal (in black) probability density functions of the tree diameter at time t = 80 year for three different values of the condition: (<b>a</b>) tree potentially available area: 6 m<sup>2</sup>—red, 26 m<sup>2</sup>—blue, 46 m<sup>2</sup>—gold; (<b>b</b>) tree height: 15 m—red, 25 m—blue, 35 m—gold; (<b>c</b>) tree crown base height: 10 m—red, 20 m—blue, 30 m—gold; (<b>d</b>) tree crown area: 5 m<sup>2</sup>—red, 25 m<sup>2</sup>—blue, 45 m<sup>2</sup>—gold.</p>
Full article ">Figure 2
<p>Relationships among the two-dimensional, marginal, and conditional entropies.</p>
Full article ">
15 pages, 500 KiB  
Article
Multivariate Extension of Raftery Copula
by Tariq Saali, Mhamed Mesfioui and Ani Shabri
Mathematics 2023, 11(2), 414; https://doi.org/10.3390/math11020414 - 12 Jan 2023
Cited by 6 | Viewed by 1147
Abstract
This paper introduces a multivariate extension of Raftery copula. The proposed copula is exchangeable and expressed in terms of order statistics. Several properties of this copula are established. In particular, the multivariate Kendall’s tau and Spearman’s rho, as well as the density function, [...] Read more.
This paper introduces a multivariate extension of Raftery copula. The proposed copula is exchangeable and expressed in terms of order statistics. Several properties of this copula are established. In particular, the multivariate Kendall’s tau and Spearman’s rho, as well as the density function, of the suggested copula are derived. The lower and upper tail dependence of the proposed copula are also established. The dependence parameter estimator of this new copula is examined based on the maximum likelihood procedure. A simulation study shows a satisfactory performance of the presented estimator. Finally, the proposed copula is successfully applied to a real data set on black cherry trees. Full article
(This article belongs to the Section Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>The curves of <math display="inline"><semantics> <msub> <mi>ρ</mi> <mn>2</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>ρ</mi> <mn>3</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>ρ</mi> <mn>4</mn> </msub> </semantics></math> in terms of <math display="inline"><semantics> <mi>θ</mi> </semantics></math> are indicated with colors red, blue and green, respectively.</p>
Full article ">Figure 2
<p>The curves of <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>4</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>5</mn> </msub> </semantics></math> in terms of <math display="inline"><semantics> <mi>θ</mi> </semantics></math> are indicated with colors blue, green and purple, respectively.</p>
Full article ">
23 pages, 4713 KiB  
Article
Modeling Number of Trees per Hectare Dynamics for Uneven-Aged, Mixed-Species Stands Using the Copula Approach
by Petras Rupšys and Edmundas Petrauskas
Forests 2023, 14(1), 12; https://doi.org/10.3390/f14010012 - 21 Dec 2022
Cited by 2 | Viewed by 1958
Abstract
For the monitoring and management of forest resources, the main index is the stand volume, which is determined on the basis of the tree diameter, height, and number of trees per hectare of three-dimensional distribution. The development of trees in the forest stand [...] Read more.
For the monitoring and management of forest resources, the main index is the stand volume, which is determined on the basis of the tree diameter, height, and number of trees per hectare of three-dimensional distribution. The development of trees in the forest stand is dynamic and is driven by random phenomena. In this study, the tree diameter, the potentially available area, and the height are described by the mixed-effect parameters of the Gompertz-type diffusion process. A normal copula function is used to connect a three-dimensional distribution to its one-dimensional margins. The newly developed model was illustrated using empirical data from 53 permanent experimental plots (measured for seven cycles), which were characterized as follows: pine forests (Pinus sylvestris), 63.8%; spruce (Picea abies), 30.2%; silver birch (Betula pendula Roth and Betula pubescens Ehrh.), 5.8%; and others, 0.2%. An analysis of the tree diameter and height of growth, including current and mean increments and inflection points, is presented. The models for the change in the number of trees per hectare with age are presented on the basis of the probabilistic density functions of the solutions of stochastic differential equations and the copula function. The dynamics of the number of trees per hectare are visualized graphically, and the goodness of fit of the newly developed models is evaluated using standard statistical measures. Full article
(This article belongs to the Special Issue Modelling Forest Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Tree potentially available area for randomly selected plot: 1983rd year cycle of measurement (mean age, 50.38 years); red—Scots pine trees; green—Norway spruce trees; yellow—birch trees; circles—tree position.</p>
Full article ">Figure 2
<p>Relationships between components of two tree sizes at a particular age: solid black line—20 years; solid blue line—60 years; solid red line—120 years; (<b>p1</b>–<b>p6</b>) pine trees; (<b>s1</b>–<b>s6</b>) spruce trees; (<b>b1</b>–<b>b6</b>) birch trees; (<b>p1</b>,<b>s1</b>,<b>b1</b>) diameter growth against potentially available area; (<b>p2</b>,<b>s2</b>,<b>b2</b>) diameter growth against height; (<b>p3</b>,<b>s3</b>,<b>b3</b>) potentially available area growth against diameter; (<b>p4</b>,<b>s4</b>,<b>b4</b>) potentially available area growth against height; (<b>p5</b>,<b>s5</b>,<b>b5</b>) height growth against diameter; (<b>p6</b>,<b>s6</b>,<b>b6</b>) height growth against potentially available area.</p>
Full article ">Figure 3
<p>Acceleration of tree diameter, potentially available area, and height growth for both the fixed- and the mixed-effect modes via age: (<b>pf1</b>–<b>pf3</b>,<b>sf1</b>–<b>sf3</b>,<b>bf1</b>–<b>bf3</b>) fixed-effect mode; (<b>pm1</b>–<b>pm3</b>,<b>sm1</b>–<b>sm3</b>,<b>bm1</b>–<b>bm3</b>) mixed-effect mode; (<b>pf1</b>–<b>pf3</b>,<b>pm1</b>–<b>pm3</b>) pine trees; (<b>sf1</b>–<b>sf3</b>,<b>sm1</b>–<b>sm3</b>) spruce trees; (<b>bf1</b>–<b>bf3</b>,<b>bm1</b>–<b>bm3</b>) birch trees; first column—diameter; second column—potentially available area; third column—height.</p>
Full article ">Figure 4
<p>Current and mean annual increments of tree diameter and height growth for both the fixed- and the mixed-effect modes via age: (<b>pf1</b>,<b>pf2</b>,<b>sf1</b>,<b>sf2</b>,<b>bf1</b>,<b>bf2</b>) fixed-effect mode; (<b>pm1</b>,<b>pm2</b>,<b>sm1</b>,<b>sm2</b>,<b>bm1</b>,<b>bm2</b>) mixed-effect mode; (<b>pf1</b>,<b>pf2</b>,<b>pm1</b>,<b>pm2</b>) pine trees; (<b>sf1</b>,<b>sf2</b>,<b>sm1</b>,<b>sm2</b>) spruce trees; (<b>bf1</b>,<b>bf2</b>,<b>bm1</b>,<b>bm2</b>) birch trees; first column—diameter; second column—height; black—first stand; blue—second stand; red—third stand; solid line—current annual increment; dotted line—mean annual increments.</p>
Full article ">Figure 4 Cont.
<p>Current and mean annual increments of tree diameter and height growth for both the fixed- and the mixed-effect modes via age: (<b>pf1</b>,<b>pf2</b>,<b>sf1</b>,<b>sf2</b>,<b>bf1</b>,<b>bf2</b>) fixed-effect mode; (<b>pm1</b>,<b>pm2</b>,<b>sm1</b>,<b>sm2</b>,<b>bm1</b>,<b>bm2</b>) mixed-effect mode; (<b>pf1</b>,<b>pf2</b>,<b>pm1</b>,<b>pm2</b>) pine trees; (<b>sf1</b>,<b>sf2</b>,<b>sm1</b>,<b>sm2</b>) spruce trees; (<b>bf1</b>,<b>bf2</b>,<b>bm1</b>,<b>bm2</b>) birch trees; first column—diameter; second column—height; black—first stand; blue—second stand; red—third stand; solid line—current annual increment; dotted line—mean annual increments.</p>
Full article ">Figure 5
<p>Relative increments multiplied by time of tree diameter, and height growth along with the signal line y = 1: (<b>pm1</b>,<b>pm2</b>) pine trees; (<b>sm1</b>,<b>sm2</b>) spruce trees; (<b>bm1</b>,<b>bm2</b>) birch trees; first column—diameter; second column—height; black—first stand; blue—second stand; red—third stand; dashed line—signal line; black—first stand; blue—second stand; red—third stand.</p>
Full article ">Figure 6
<p>Linkage of current increments of tree diameter and height growth with mean diameter and height: (<b>p1</b>,<b>p2</b>) pine trees; (<b>s1</b>,<b>s2</b>) spruce trees; (<b>b1</b>,<b>b2</b>) birch trees; first column—diameter; second column—height; black—first stand; blue—second stand; red—third stand.</p>
Full article ">Figure 7
<p>Evolution of the number of trees per hectare for three randomly selected stands: black—first stand; blue—second stand; red—third stand; (<b>a</b>) all trees; (<b>b</b>) pine trees (<math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>p</mi> <mi>i</mi> <mi>n</mi> <mi>e</mi> </mrow> <mn>1</mn> </msubsup> <mo>=</mo> <mn>0.97</mn> </mrow> </semantics></math> in black, <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>p</mi> <mi>i</mi> <mi>n</mi> <mi>e</mi> </mrow> <mn>2</mn> </msubsup> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math> in blue; <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>p</mi> <mi>i</mi> <mi>n</mi> <mi>e</mi> </mrow> <mn>3</mn> </msubsup> <mo>=</mo> <mn>0.73</mn> </mrow> </semantics></math> in red); (<b>c</b>) spruce trees (<math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>s</mi> <mi>p</mi> <mi>r</mi> <mi>u</mi> <mi>c</mi> <mi>e</mi> </mrow> <mn>1</mn> </msubsup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math> in black, <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>s</mi> <mi>p</mi> <mi>r</mi> <mi>u</mi> <mi>c</mi> <mi>e</mi> </mrow> <mn>2</mn> </msubsup> <mo>=</mo> <mn>0.72</mn> </mrow> </semantics></math> in blue; <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>s</mi> <mi>p</mi> <mi>r</mi> <mi>u</mi> <mi>c</mi> <mi>e</mi> </mrow> <mn>3</mn> </msubsup> <mo>=</mo> <mn>0.31</mn> </mrow> </semantics></math> in red); (<b>d</b>) birch tree (<math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>r</mi> <mi>c</mi> <mi>h</mi> </mrow> <mn>1</mn> </msubsup> <mo>=</mo> <mn>0.22</mn> </mrow> </semantics></math> in black, <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>r</mi> <mi>c</mi> <mi>h</mi> </mrow> <mn>2</mn> </msubsup> <mo>=</mo> <mn>0.16</mn> </mrow> </semantics></math> in blue; <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>r</mi> <mi>c</mi> <mi>h</mi> </mrow> <mn>3</mn> </msubsup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math> in red).</p>
Full article ">Figure 8
<p>Evolution of the relative annual decay process of the number of trees per hectare for three randomly selected stands: black—first stand; blue—second stand; red—third stand; (<b>a</b>) all trees; (<b>b</b>) pine trees (<math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>p</mi> <mi>i</mi> <mi>n</mi> <mi>e</mi> </mrow> <mn>1</mn> </msubsup> <mo>=</mo> <mn>0.97</mn> </mrow> </semantics></math> in black, <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>p</mi> <mi>i</mi> <mi>n</mi> <mi>e</mi> </mrow> <mn>2</mn> </msubsup> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math> in blue; <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>p</mi> <mi>i</mi> <mi>n</mi> <mi>e</mi> </mrow> <mn>3</mn> </msubsup> <mo>=</mo> <mn>0.73</mn> </mrow> </semantics></math> in red); (<b>c</b>) spruce trees (<math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>s</mi> <mi>p</mi> <mi>r</mi> <mi>u</mi> <mi>c</mi> <mi>e</mi> </mrow> <mn>1</mn> </msubsup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math> in black, <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>s</mi> <mi>p</mi> <mi>r</mi> <mi>u</mi> <mi>c</mi> <mi>e</mi> </mrow> <mn>2</mn> </msubsup> <mo>=</mo> <mn>0.72</mn> </mrow> </semantics></math> in blue; <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>s</mi> <mi>p</mi> <mi>r</mi> <mi>u</mi> <mi>c</mi> <mi>e</mi> </mrow> <mn>3</mn> </msubsup> <mo>=</mo> <mn>0.31</mn> </mrow> </semantics></math> in red); (<b>d</b>) birch tree (<math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>r</mi> <mi>c</mi> <mi>h</mi> </mrow> <mn>1</mn> </msubsup> <mo>=</mo> <mn>0.22</mn> </mrow> </semantics></math> in black, <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>r</mi> <mi>c</mi> <mi>h</mi> </mrow> <mn>2</mn> </msubsup> <mo>=</mo> <mn>0.16</mn> </mrow> </semantics></math> in blue; <math display="inline"><semantics> <mrow> <msubsup> <mi>k</mi> <mrow> <mi>b</mi> <mi>i</mi> <mi>r</mi> <mi>c</mi> <mi>h</mi> </mrow> <mn>3</mn> </msubsup> <mo>=</mo> <mn>0.05</mn> </mrow> </semantics></math> in red).</p>
Full article ">
20 pages, 7477 KiB  
Article
On the Construction of Growth Models via Symmetric Copulas and Stochastic Differential Equations
by Petras Rupšys and Edmundas Petrauskas
Symmetry 2022, 14(10), 2127; https://doi.org/10.3390/sym14102127 - 12 Oct 2022
Cited by 5 | Viewed by 1594
Abstract
By nature, growth regulatory networks in biology are dynamic and stochastic, and feedback regulates their growth function at different ages. In this study, we carried out a stochastic modeling of growth networks and demonstrated this method using three mixed effect four-parameter Gompertz-type diffusion [...] Read more.
By nature, growth regulatory networks in biology are dynamic and stochastic, and feedback regulates their growth function at different ages. In this study, we carried out a stochastic modeling of growth networks and demonstrated this method using three mixed effect four-parameter Gompertz-type diffusion processes and a combination thereof using the conditional normal copula function. Using the conditional normal copula, newly derived univariate distributions can be combined into trivariate and bivariate distributions, and their corresponding conditional bivariate and univariate distributions. The link between the predictor variable and the remaining one or two explanatory variables can be formalized using copula-type densities and a numerical integration procedure. In this study, for parameter estimation, we used a semiparametric maximum pseudo-likelihood estimator procedure, which was characterized by a two-step technique, namely, separately estimating the parameters of the marginal distributions and the parameters of the copula. The results were illustrated using two observed longitudinal datasets, the first of which included the age, diameter, and potentially available area of 39,437 trees (48 stands), while the second included the age, diameter, potentially available area, and height of 8604 trees (47 stands) covering uneven mixed-species (pine, spruce, and birch) stands. All results were implemented using the MAPLE symbolic algebra system. Full article
(This article belongs to the Section Mathematics)
Show Figures

Figure 1

Figure 1
<p>Bivariate conditional densities for the fixed effect scenario: (<b>a1</b>) height of 15 m, time 60 years; (<b>a2</b>) height of 35 m, time 60 years; (<b>b1</b>) area of 10 m<sup>2</sup>, time 60 years; (<b>b2</b>) area of 50 m<sup>2</sup>, time 60 years; (<b>c1</b>) diameter of 10 cm, time 60 years; (<b>c2</b>) diameter of 30 cm, time 60 years.</p>
Full article ">Figure 1 Cont.
<p>Bivariate conditional densities for the fixed effect scenario: (<b>a1</b>) height of 15 m, time 60 years; (<b>a2</b>) height of 35 m, time 60 years; (<b>b1</b>) area of 10 m<sup>2</sup>, time 60 years; (<b>b2</b>) area of 50 m<sup>2</sup>, time 60 years; (<b>c1</b>) diameter of 10 cm, time 60 years; (<b>c2</b>) diameter of 30 cm, time 60 years.</p>
Full article ">Figure 2
<p>Bivariate conditional densities for the mixed effect scenario and one randomly selected stand: (<b>a1</b>) height of 15 m, time 60 years; (<b>a2</b>) height of 35 m, time 60 years; (<b>b1</b>) area of 10 m<sup>2</sup>, time 60 years; (<b>b2</b>) area of 50 m<sup>2</sup>, time 60 years; (<b>c1</b>) diameter of 10 cm, time 60 years; (<b>c2</b>) diameter of 30 cm, time 60 years.</p>
Full article ">Figure 2 Cont.
<p>Bivariate conditional densities for the mixed effect scenario and one randomly selected stand: (<b>a1</b>) height of 15 m, time 60 years; (<b>a2</b>) height of 35 m, time 60 years; (<b>b1</b>) area of 10 m<sup>2</sup>, time 60 years; (<b>b2</b>) area of 50 m<sup>2</sup>, time 60 years; (<b>c1</b>) diameter of 10 cm, time 60 years; (<b>c2</b>) diameter of 30 cm, time 60 years.</p>
Full article ">Figure 3
<p>Bivariate copula-type densities and contour lines for the mixed effect scenario: (<b>a1</b>) bivariate density of diameter and area, time 30 years; (<b>a2</b>) bivariate density of diameter and area, time 80 years; (<b>a3</b>) contour line of diameter and area, 30 years; (<b>a4</b>) contour line of diameter and area, time 60 years; (<b>b1</b>) bivariate density of diameter and height, time 30 years; (<b>b2</b>) bivariate density of diameter and height, time 80 years; (<b>b3</b>) contour line of diameter and height, time 30 years; (<b>b4</b>) contour line of diameter and height, time 60 years; (<b>c1</b>) bivariate density of area and height, time 30 years; (<b>c2</b>) bivariate density of area and height, time 80 years; (<b>c3</b>) contour line of area and height, time 30 years; (<b>c4</b>) contour line of area and height, time 60 years. Observed dataset shown in boxes.</p>
Full article ">Figure 3 Cont.
<p>Bivariate copula-type densities and contour lines for the mixed effect scenario: (<b>a1</b>) bivariate density of diameter and area, time 30 years; (<b>a2</b>) bivariate density of diameter and area, time 80 years; (<b>a3</b>) contour line of diameter and area, 30 years; (<b>a4</b>) contour line of diameter and area, time 60 years; (<b>b1</b>) bivariate density of diameter and height, time 30 years; (<b>b2</b>) bivariate density of diameter and height, time 80 years; (<b>b3</b>) contour line of diameter and height, time 30 years; (<b>b4</b>) contour line of diameter and height, time 60 years; (<b>c1</b>) bivariate density of area and height, time 30 years; (<b>c2</b>) bivariate density of area and height, time 80 years; (<b>c3</b>) contour line of area and height, time 30 years; (<b>c4</b>) contour line of area and height, time 60 years. Observed dataset shown in boxes.</p>
Full article ">Figure 4
<p>Bivariate copula-type densities and contour lines for the mixed effect scenario and two randomly selected stands: (<b>a1</b>) first stand, bivariate density of diameter and area, time 60 years; (<b>a2</b>) second stand, bivariate density of diameter and area, time 60 years; (<b>a3</b>) first stand, contour line of diameter and area, 60 years; (<b>a4</b>) second stand, contour line of diameter and area, time 60 years; (<b>b1</b>) first stand, bivariate density of diameter and height, time 60 years; (<b>b2</b>) second stand, bivariate density of diameter and height, time 60 years; (<b>b3</b>) first stand, contour line of diameter and height, 60 years; (<b>b4</b>) second stand, contour line of diameter and height, time 60 years; (<b>c1</b>) first stand, bivariate density of area and height, time 60 years; (<b>c2</b>) second stand, bivariate density of area and height, time 60 years; (<b>c3</b>) first stand, contour line of area and height, time 60 years; (<b>c4</b>) second stand, contour line of area and height, time 60 years. Observed dataset shown in boxes.</p>
Full article ">Figure 4 Cont.
<p>Bivariate copula-type densities and contour lines for the mixed effect scenario and two randomly selected stands: (<b>a1</b>) first stand, bivariate density of diameter and area, time 60 years; (<b>a2</b>) second stand, bivariate density of diameter and area, time 60 years; (<b>a3</b>) first stand, contour line of diameter and area, 60 years; (<b>a4</b>) second stand, contour line of diameter and area, time 60 years; (<b>b1</b>) first stand, bivariate density of diameter and height, time 60 years; (<b>b2</b>) second stand, bivariate density of diameter and height, time 60 years; (<b>b3</b>) first stand, contour line of diameter and height, 60 years; (<b>b4</b>) second stand, contour line of diameter and height, time 60 years; (<b>c1</b>) first stand, bivariate density of area and height, time 60 years; (<b>c2</b>) second stand, bivariate density of area and height, time 60 years; (<b>c3</b>) first stand, contour line of area and height, time 60 years; (<b>c4</b>) second stand, contour line of area and height, time 60 years. Observed dataset shown in boxes.</p>
Full article ">Figure 5
<p>Univariate conditional distributions of copula-based densities at age 60 under the fixed effect scenario: (<b>a1</b>) univariate conditional distribution of diameter from area; (<b>a2</b>) univariate conditional distribution of diameter from height; (<b>b1</b>) univariate conditional distribution of area from diameter; (<b>b2</b>) univariate conditional distribution of area from height; (<b>c1</b>) univariate conditional distribution of height from diameter; (<b>c2</b>) univariate conditional distribution of height from area.</p>
Full article ">Figure 6
<p>Univariate conditional distributions of copula-based densities at age 60 under the mixed effect scenario: (<b>a1</b>) univariate conditional distribution of diameter from area; (<b>a2</b>) univariate conditional distribution of diameter from height; (<b>b1</b>) univariate conditional distribution of area from diameter; (<b>b2</b>) univariate conditional distribution of area from height; (<b>c1</b>) univariate conditional distribution of height from diameter; (<b>c2</b>) univariate conditional distribution of height from area.</p>
Full article ">Figure 7
<p>Trends of the diameter, potentially available area, and height for two randomly selected stands under the mixed effect scenario: (<b>a1</b>) first stand, diameter trends; (<b>a2</b>) second stand, diameter trends; (<b>b1</b>) first stand, potentially available area trends; (<b>b2</b>) second stand, potentially available area trends; (<b>c1</b>) first stand, height trends; (<b>c2</b>) second stand, height trends. Solid black line—mean trend; solid red line—median trend; solid blue line—mode trend; dashed black lines—quantile trends. Observed data shown in circles.</p>
Full article ">Figure 7 Cont.
<p>Trends of the diameter, potentially available area, and height for two randomly selected stands under the mixed effect scenario: (<b>a1</b>) first stand, diameter trends; (<b>a2</b>) second stand, diameter trends; (<b>b1</b>) first stand, potentially available area trends; (<b>b2</b>) second stand, potentially available area trends; (<b>c1</b>) first stand, height trends; (<b>c2</b>) second stand, height trends. Solid black line—mean trend; solid red line—median trend; solid blue line—mode trend; dashed black lines—quantile trends. Observed data shown in circles.</p>
Full article ">
24 pages, 1278 KiB  
Article
Bivariate Copula Trees for Gross Loss Aggregation with Positively Dependent Risks
by Rafał Wójcik and Charlie Wusuo Liu
Risks 2022, 10(8), 144; https://doi.org/10.3390/risks10080144 - 22 Jul 2022
Viewed by 3373
Abstract
We propose several numerical algorithms to compute the distribution of gross loss in a positively dependent catastrophe insurance portfolio. Hierarchical risk aggregation is performed using bivariate copula trees. Six common parametric copula families are studied. At every branching node, the distribution of a [...] Read more.
We propose several numerical algorithms to compute the distribution of gross loss in a positively dependent catastrophe insurance portfolio. Hierarchical risk aggregation is performed using bivariate copula trees. Six common parametric copula families are studied. At every branching node, the distribution of a sum of risks is obtained by discrete copula convolution. This approach is compared to approximation by a weighted average of independent and comonotonic distributions. The weight is a measure of positive dependence through variance of the aggregate risk. During gross loss accumulation, the marginals are distorted by application of insurance financial terms, and the value of the mixing weight is impacted. To accelerate computations, we capture this effect using the ratio of standard deviations of pre-term and post-term risks, followed by covariance scaling. We test the performance of our algorithms using three examples of complex insurance portfolios subject to hurricane and earthquake catastrophes. Full article
Show Figures

Figure 1

Figure 1
<p>Examples of typical supermodular functions used in catastrophe insurance loss aggregation. Red surface represents <math display="inline"><semantics> <mrow> <mi>g</mi> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>+</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> </mrow> </semantics></math>. Plotted in blue are (<b>A</b>) <math display="inline"><semantics> <mrow> <mi>g</mi> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>+</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>,</mo> <mn>3</mn> <mo>)</mo> </mrow> <mo>+</mo> </msub> </mrow> </semantics></math>, (<b>B</b>) <math display="inline"><semantics> <mrow> <mi>g</mi> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <msub> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>−</mo> <mn>3</mn> <mo>)</mo> </mrow> <mo>+</mo> </msub> <mo>,</mo> <mn>5</mn> <mo>)</mo> </mrow> <mo>+</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <msub> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>−</mo> <mn>3</mn> <mo>)</mo> </mrow> <mo>+</mo> </msub> <mo>,</mo> <mn>5</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and (<b>C</b>) <math display="inline"><semantics> <mrow> <mi>g</mi> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <msub> <mrow> <mo>(</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>1</mn> </msub> <mo>,</mo> <mn>3</mn> <mo>)</mo> </mrow> <mo>+</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <msub> <mi>x</mi> <mn>2</mn> </msub> <mo>,</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>,</mo> <mn>2</mn> <mo>)</mo> </mrow> <mo>+</mo> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Computing the sum of three risks using direct aggregation tree (<b>A</b>–<b>C</b>) and hierarchical aggregation tree with sequential topology (<b>D</b>–<b>G</b>) from ground-up loss perspective (left column) and gross loss perspective (middle and right columns). The branching nodes of hierarchical trees (black dots) represent summation of the incoming pairs of individual and/or cumulative risks. For gross loss perspective, transformation nodes (white dots) represent application of the financial terms <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>,</mo> <mi>ψ</mi> </mrow> </semantics></math> to individual and/or cumulative risks.</p>
Full article ">Figure 3
<p>An example of stop-loss order preservation under truncation transform. Here, <span class="html-italic">X</span> and <span class="html-italic">Y</span> are random variables with discrete marginals <math display="inline"><semantics> <msub> <mi>p</mi> <mi>X</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>p</mi> <mi>Y</mi> </msub> </semantics></math> in (<b>A</b>,<b>B</b>), respectively. The distribution <math display="inline"><semantics> <msub> <mi>p</mi> <msup> <mi>S</mi> <mo>⊥</mo> </msup> </msub> </semantics></math> of the independent sum <math display="inline"><semantics> <mrow> <msup> <mi>S</mi> <mo>⊥</mo> </msup> <mo>=</mo> <msup> <mi>X</mi> <mo>⊥</mo> </msup> <mo>+</mo> <msup> <mi>Y</mi> <mo>⊥</mo> </msup> </mrow> </semantics></math> in (<b>C</b>) is obtained by discrete convolution (Algorithm A1 in <a href="#B50-risks-10-00144" class="html-bibr">Wójcik et al. 2019</a>), while the distribution <math display="inline"><semantics> <msub> <mi>p</mi> <msup> <mi>S</mi> <mo>+</mo> </msup> </msub> </semantics></math> of the comonotonic sum <math display="inline"><semantics> <mrow> <msup> <mi>S</mi> <mo>+</mo> </msup> <mo>=</mo> <msup> <mi>X</mi> <mo>+</mo> </msup> <mo>+</mo> <msup> <mi>Y</mi> <mo>+</mo> </msup> </mrow> </semantics></math> in (<b>D</b>) is computed using numerical quantile addition (Algorithm 6 in <a href="#B50-risks-10-00144" class="html-bibr">Wójcik et al. 2019</a>). The dashed vertical lines represent the truncation transform bounds <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1.75</mn> </mrow> </semantics></math>. In (<b>E</b>), the alignment of the corresponding cdfs <math display="inline"><semantics> <msub> <mi>P</mi> <msup> <mi>S</mi> <mo>⊥</mo> </msup> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>P</mi> <msup> <mi>S</mi> <mo>+</mo> </msup> </msub> </semantics></math> shows that <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <msup> <mi>S</mi> <mo>⊥</mo> </msup> </msub> <mrow> <mo>(</mo> <mi>b</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <msub> <mi>P</mi> <msup> <mi>S</mi> <mo>+</mo> </msup> </msub> <mrow> <mo>(</mo> <mi>b</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1.75</mn> </mrow> </semantics></math>, so (<a href="#FD29-risks-10-00144" class="html-disp-formula">29</a>) holds. This, together with the necessary condition (<a href="#FD28-risks-10-00144" class="html-disp-formula">28</a>), implies that the truncated cdfs in (<b>F</b>,<b>G</b>) characterize the stop-loss order <math display="inline"><semantics> <mrow> <msup> <mi>S</mi> <mo>⊥</mo> </msup> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> </mrow> <msub> <mo>≤</mo> <mrow> <mi>s</mi> <mi>l</mi> </mrow> </msub> <msup> <mi>S</mi> <mo>+</mo> </msup> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>. In (<b>H</b>), the binary decision whether the necessary condition is true or false is plotted as a function of the truncation bounds <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>a</mi> <mo>&lt;</mo> <mi>b</mi> <mo>≤</mo> <mn>2</mn> </mrow> </semantics></math>. The means <math display="inline"><semantics> <msub> <mi>μ</mi> <mrow> <msup> <mi>S</mi> <mo>⊥</mo> </msup> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> </mrow> </mrow> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>μ</mi> <mrow> <msup> <mi>S</mi> <mo>+</mo> </msup> <mrow> <mo>(</mo> <mi>a</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> </mrow> </mrow> </msub> </semantics></math> characterize the independent and comonotonic sums after application of the truncation transform. The red region is where the necessary condition holds. The black dot represents the actual truncation bounds used throughout this example. The impermissible region where <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>&lt;</mo> <mi>a</mi> </mrow> </semantics></math> is plotted in grey.</p>
Full article ">Figure 4
<p>Three generic summation nodes: (<b>A</b>) ground-up node, (<b>B</b>) gross loss node and (<b>C</b>) back-allocated version of the gross loss node.</p>
Full article ">Figure 5
<p>Special cases of copula distributions: (<b>A</b>) the independence copula <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>(</mo> <mi>u</mi> <mo>,</mo> <mi>v</mi> <mo>)</mo> <mo>=</mo> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math>, (<b>B</b>) the comonotonicity copula <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>(</mo> <mi>u</mi> <mo>,</mo> <mi>v</mi> <mo>)</mo> <mo>=</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <mi>u</mi> <mo>,</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and (<b>C</b>) Fréchet copula <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>(</mo> <mi>u</mi> <mo>,</mo> <mi>v</mi> <mo>)</mo> <mo>=</mo> <mo>(</mo> <mn>1</mn> <mo>−</mo> <mi>w</mi> <mo>)</mo> <mspace width="0.166667em"/> <mi>u</mi> <mi>v</mi> <mo>+</mo> <mi>w</mi> <mspace width="0.166667em"/> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <mi>u</mi> <mo>,</mo> <mi>v</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Illustration of the copula decomposition: (<b>A</b>) Joe copula in <a href="#risks-10-00144-t001" class="html-table">Table 1</a> with <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mn>0.07</mn> </mrow> </semantics></math> discretized on 11 × 11 grid, (<b>B</b>) Fréchet decomposition of Joe copula, (<b>C</b>) the bivariate pmf <math display="inline"><semantics> <msub> <mi>p</mi> <mrow> <mi>X</mi> <mo>,</mo> <mi>Y</mi> </mrow> </msub> </semantics></math> obtained by combining the discretized <math display="inline"><semantics> <mrow> <mi>G</mi> <mi>a</mi> <mi>m</mi> <mi>m</mi> <mi>a</mi> <mo>(</mo> <mn>5</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math> marginals <math display="inline"><semantics> <msub> <mi>p</mi> <mi>X</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>p</mi> <mi>Y</mi> </msub> </semantics></math> (black bars) using the discretized Joe copula and (<b>D</b>) Fréchet decomposition of <math display="inline"><semantics> <msub> <mi>p</mi> <mrow> <mi>X</mi> <mo>,</mo> <mi>Y</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 7
<p>Actual (relative) execution time in nanoseconds for computing <math display="inline"><semantics> <msub> <mi>p</mi> <mrow> <msub> <mi>ϕ</mi> <mi>X</mi> </msub> <mrow> <mo>(</mo> <mi>X</mi> <mo>)</mo> </mrow> <mo>+</mo> <msub> <mi>ϕ</mi> <mi>Y</mi> </msub> <mrow> <mo>(</mo> <mi>Y</mi> <mo>)</mo> </mrow> </mrow> </msub> </semantics></math>. Here, the support of <span class="html-italic">X</span> is <math display="inline"><semantics> <mrow> <mi>Ran</mi> <mspace width="0.277778em"/> <mi>X</mi> <mo>=</mo> <mo>{</mo> <mn>0</mn> <mo>,</mo> <mn>0.1429</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.2857</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.4286</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.5714</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.7143</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.8571</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>1</mn> <mo>}</mo> </mrow> </semantics></math>, and the support of <span class="html-italic">Y</span> is <math display="inline"><semantics> <mrow> <mi>Ran</mi> <mspace width="0.277778em"/> <mi>Y</mi> <mo>=</mo> <mo>{</mo> <mn>0</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.1429</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.2857</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.4286</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.5714</mn> <mo>,</mo> <mn>0.7143</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.8571</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>1</mn> <mo>}</mo> </mrow> </semantics></math>, with probabilities <math display="inline"><semantics> <mrow> <mi>Ran</mi> <mspace width="0.277778em"/> <msub> <mi>p</mi> <mi>X</mi> </msub> <mrow> <mo>=</mo> <mo>{</mo> <mn>0.2327</mn> <mo>,</mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0268</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0051</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0493</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.3023</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.1834</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0093</mn> <mo>,</mo> <mn>0.1911</mn> <mo>}</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Ran</mi> <mspace width="0.277778em"/> <msub> <mi>p</mi> <mi>Y</mi> </msub> <mrow> <mo>=</mo> <mo>{</mo> <mn>0.1730</mn> <mo>,</mo> </mrow> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0666</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.3864</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.1648</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0021</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0703</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0871</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.0497</mn> <mo>}</mo> </mrow> </semantics></math>. Correlation values are <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>(</mo> <mi>X</mi> <mo>,</mo> <mi>Y</mi> <mo>)</mo> <mo>=</mo> <mo>{</mo> <mn>0</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.01</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mn>0.02</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mo>…</mo> <mo>,</mo> <mn>1</mn> <mo>}</mo> </mrow> </semantics></math> and the financial terms <math display="inline"><semantics> <mrow> <msub> <mi>ϕ</mi> <mi>X</mi> </msub> <mrow> <mo>(</mo> <mi>X</mi> <mo>)</mo> </mrow> <mo>=</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <msub> <mrow> <mo>(</mo> <mi>X</mi> <mo>−</mo> <mn>0.2</mn> <mo>)</mo> </mrow> <mo>+</mo> </msub> <mo>,</mo> <mn>0.9</mn> <mo>)</mo> </mrow> <mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <msub> <mi>ϕ</mi> <mi>Y</mi> </msub> <mrow> <mo>(</mo> <mi>Y</mi> <mo>)</mo> </mrow> <mo>=</mo> <mo movablelimits="true" form="prefix">min</mo> <mrow> <mo>(</mo> <msub> <mrow> <mo>(</mo> <mi>Y</mi> <mo>−</mo> <mn>0.1</mn> <mo>)</mo> </mrow> <mo>+</mo> </msub> <mo>,</mo> <mn>0.8</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. Computing time is the average over 101 runs. Platform: Intel i9-9980HK CPU, 32 GB RAM, Windows 10. Compiler: Mingw-w64 g++ 8.3 -std=gnu++17 -Ofast -mfpmath=sse -msse2 -mstackrealign.</p>
Full article ">Figure 8
<p>Three examples of marginal damage distributions. The mean damage ratio varies from low in (<b>A</b>) to moderate in (<b>B</b>) and high in (<b>C</b>).</p>
Full article ">Figure 9
<p>Total gross loss pmfs for Portfolio 1. The pmfs are plotted within the same x-axis range for clarity. Blue pmfs are obtained using aggregation tree with copulas in <a href="#risks-10-00144-t001" class="html-table">Table 1</a> at summation nodes. Red pmfs are obtained by replacing each copula with its corresponding Fréchet decomposition into comonotonic part and independent part.</p>
Full article ">Figure 10
<p>Total gross loss pmfs for Portfolio 2. The pmfs are plotted within the same x-axis range for clarity. Blue pmfs are obtained using aggregation tree with copulas in <a href="#risks-10-00144-t001" class="html-table">Table 1</a> at summation nodes. Red pmfs are obtained by replacing each copula with its corresponding Fréchet decomposition into comonotonic part and independent part.</p>
Full article ">Figure 11
<p>Total gross loss pmfs for Portfolio 3. The pmfs are plotted within the same x-axis range for clarity. Blue pmfs are obtained using aggregation tree with copulas in <a href="#risks-10-00144-t001" class="html-table">Table 1</a> at summation nodes. Red pmfs are obtained by replacing each copula with its corresponding Fréchet decomposition into comonotonic part and independent part.</p>
Full article ">
18 pages, 371 KiB  
Article
Application of Deep Learning and Neural Network to Speeding Ticket and Insurance Claim Count Data
by Jong-Min Kim, Jihun Kim and Il Do Ha
Axioms 2022, 11(6), 280; https://doi.org/10.3390/axioms11060280 - 10 Jun 2022
Cited by 1 | Viewed by 2995
Abstract
With the popularity of big data analysis with insurance claim count data, diverse regression models for count response variable have been developed. However, there is a multicollinearlity issue with multivariate input variables to the count response regression models. Recently, deep learning and neural [...] Read more.
With the popularity of big data analysis with insurance claim count data, diverse regression models for count response variable have been developed. However, there is a multicollinearlity issue with multivariate input variables to the count response regression models. Recently, deep learning and neural network models for count response have been proposed, and a Keras and Tensorflow-based deep learning model has been also proposed. To apply the deep learning and neural network models to non-normal insurance claim count data, we perform the root mean square error accuracy comparison of gradient boosting machines (a popular machine learning regression tree algorithm), principal component analysis (PCA)-based Poisson regression, PCA-based negative binomial regression, and PCA-based zero inflated poisson regression to avoid the multicollinearity of multivariate input variables with the simulated normal distribution data and the non-normal simulated data combined with normally distributed data, binary data, copula-based asymmetrical data, and two real data sets, which consist of speeding ticket and Singapore insurance claim count data. Full article
(This article belongs to the Special Issue Machine Learning: Theory, Algorithms and Applications)
Show Figures

Figure 1

Figure 1
<p>A feed-forward deep learning model with two-hidden layers where bias terms are omitted for brevity but are written in the main text.</p>
Full article ">Figure 2
<p>Violin Plots of RMSE with Simulated Multivariate Normal Data and Non-Normal Combined Data with Multivariate Normal, Copula and Binary Data.</p>
Full article ">Figure 3
<p>Plots of RMSE with real data.</p>
Full article ">
17 pages, 6517 KiB  
Article
Dynamic Time Warping Algorithm in Modeling Systemic Risk in the European Insurance Sector
by Anna Denkowska and Stanisław Wanat
Entropy 2021, 23(8), 1022; https://doi.org/10.3390/e23081022 - 8 Aug 2021
Cited by 8 | Viewed by 2427
Abstract
We are looking for tools to identify, model, and measure systemic risk in the insurance sector. To this aim, we investigated the possibilities of using the Dynamic Time Warping (DTW) algorithm in two ways. The first way of using DTW is to assess [...] Read more.
We are looking for tools to identify, model, and measure systemic risk in the insurance sector. To this aim, we investigated the possibilities of using the Dynamic Time Warping (DTW) algorithm in two ways. The first way of using DTW is to assess the suitability of the Minimum Spanning Trees’ (MST) topological indicators, which were constructed based on the tail dependence coefficients determined by the copula-DCC-GARCH model in order to establish the links between insurance companies in the context of potential shock contagion. The second way consists of using the DTW algorithm to group institutions by the similarity of their contribution to systemic risk, as expressed by DeltaCoVaR, in the periods distinguished. For the crises and the normal states identified during the period 2005–2019 in Europe, we analyzed the similarity of the time series of the topological indicators of MST, constructed for 38 European insurance institutions. The results obtained confirm the effectiveness of MST topological indicators for systemic risk identification and the evaluation of indirect links between insurance institutions. Full article
(This article belongs to the Special Issue Fractal and Multifractal Analysis of Complex Networks)
Show Figures

Figure 1

Figure 1
<p>MST’s topological indicators in the periods under study. Own source.</p>
Full article ">Figure 2
<p>Average Path Length in the periods under study. Own source.</p>
Full article ">Figure 3
<p>Dendrogram for APL-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 4
<p>Diameter in the periods under study. Own source.</p>
Full article ">Figure 5
<p>Dendrogram for Diameter-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 6
<p>Max.Degree in the periods under study. Own source.</p>
Full article ">Figure 7
<p>Dendrogram for Max.Degree-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 8
<p>Alpha in the periods under study. Own source.</p>
Full article ">Figure 9
<p>Dendrogram for alpha-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 10
<p>RCE in the periods under study. Own source.</p>
Full article ">Figure 11
<p>Dendrogram for RCE-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 12
<p>Assortativity in the periods under study. Own source.</p>
Full article ">Figure 13
<p>Dendrogram for assortativity-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 14
<p>DeltaCoVaR measure for the analyzed market states. Own source.</p>
Full article ">Figure 15
<p>Average DeltaCoVaR for all analyzed institutions in the periods under study. Own source.</p>
Full article ">Figure 16
<p>Dendrogram for DeltaCoVaR-based hierarchical clustering of the periods under study. Own source.</p>
Full article ">Figure 17
<p>MST based on DTW in Normal state. Own source.</p>
Full article ">Figure 18
<p>MST based on DTW in SMC state. Own source.</p>
Full article ">Figure 19
<p>MST based on DTW in I state. Own source.</p>
Full article ">Figure 20
<p>MST based on DTW in FIC state. Own source.</p>
Full article ">
16 pages, 3833 KiB  
Article
Vine-Copula-Based Quantile Regression for Cascade Reservoirs Management
by Wafaa El Hannoun, Salah-Eddine El Adlouni and Abdelhak Zoglat
Water 2021, 13(7), 964; https://doi.org/10.3390/w13070964 - 31 Mar 2021
Cited by 4 | Viewed by 2778
Abstract
This paper features an application of Regular Vine (R-vine) copulas, a recently developed statistical tool to assess composite risk. Copula-based dependence modelling is a popular tool in conditional risk assessment, but is usually applied to pairs of variables. By contrast, Vine copulas provide [...] Read more.
This paper features an application of Regular Vine (R-vine) copulas, a recently developed statistical tool to assess composite risk. Copula-based dependence modelling is a popular tool in conditional risk assessment, but is usually applied to pairs of variables. By contrast, Vine copulas provide greater flexibility and permit the modelling of complex dependency patterns using a wide variety of bivariate copulas which may be arranged and analysed in a tree structure to explore multiple dependencies. This study emphasises the use of R-vine copulas in an analysis of the co-dependencies of five reservoirs in the cascade of the Saint-John River basin in Eastern Canada. The developed R-vine copulas lead to the joint and conditional return periods of maximum volumes, for hydrologic design and cascade reservoir management in the basin. The main attraction of this approach to risk modelling is the flexibility in the choice of distributions used to model heavy-tailed marginals and co-dependencies. Full article
(This article belongs to the Special Issue Statistical Approach to Hydrological Analysis)
Show Figures

Figure 1

Figure 1
<p>Three different vine structures obtained from the Saint-John basin structure. The blue circles represent the vine nodes which are the analyzed control stations listed in <a href="#water-13-00964-t001" class="html-table">Table 1</a>. The thick black arrows represent the vine edges. The black lines represent the sub-catchment delimitations. (<b>a</b>) A C-vine with 4 nodes. (<b>b</b>) A D-vine with 4 nodes. (<b>c</b>) An R-vine with 5 nodes.</p>
Full article ">Figure 2
<p>The Saint-John basin and the studied sub-basins.</p>
Full article ">Figure 3
<p>Location of the control sections used to illustrate the Vine-copula approach.</p>
Full article ">Figure 4
<p>The reduced-network R-vine. V1, V2, V3, V4 and V5 stand for stations 3, 7, 8, 11 and 15, respectively. (<b>a</b>) Tree 1. (<b>b</b>) Tree 2. (<b>c</b>) Tree 3. (<b>d</b>) Tree 4.</p>
Full article ">Figure 5
<p>Fitted R-vine normalized contour plots of the specified pair copulas for the volume variable.</p>
Full article ">Figure 6
<p>Pair-plots (bottom left of the diagonal), Kendall’s <math display="inline"><semantics> <mi>τ</mi> </semantics></math> values (top right of the diagonal), and the histograms (diagonal) of the volume variable in the five stations.</p>
Full article ">Figure 7
<p>Tau D-vine of the volume variable. (<b>a</b>) Tree 1. (<b>b</b>) Tree 2. (<b>c</b>) Tree 3. (<b>d</b>) Tree 4.</p>
Full article ">Figure 8
<p>Illustration of the obtained scenarios with respect to downstream conditional quantiles. The red color means values are set to <math display="inline"><semantics> <msub> <mi>Q</mi> <mn>100</mn> </msub> </semantics></math> as in <a href="#water-13-00964-t005" class="html-table">Table 5</a>. White means values are set to <math display="inline"><semantics> <msub> <mi>Q</mi> <mn>2</mn> </msub> </semantics></math>. For station 15, blue means the conditional quantile S15 <math display="inline"><semantics> <msub> <mi>Q</mi> <mn>100</mn> </msub> </semantics></math> is less than the marginal index, it is pink when they are almost equal and red when S15 <math display="inline"><semantics> <msub> <mi>Q</mi> <mn>100</mn> </msub> </semantics></math> is much greater than the marginal index. (<b>a–f</b>): scenarios 5, 8, 2, 6, 1, and 3 respectively.</p>
Full article ">Figure 9
<p>Conditional quantiles of the volume (1000 m<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>) in station 15 according to the variations in probabilities with fixed covariates.</p>
Full article ">Figure A1
<p>Time series, histograms with densities of theoretical laws and QQ-plots of the volume variable of the five studied stations. From top to bottom: station 3, 7, 8, 11 and 15.</p>
Full article ">
16 pages, 489 KiB  
Article
Copula-Based Bayesian Reliability Analysis of a Product of a Probability and a Frequency Model for Parallel Systems When Components Are Dependent
by Shi-Woei Lin, Tapiwa Blessing Matanhire and Yi-Ting Liu
Appl. Sci. 2021, 11(4), 1697; https://doi.org/10.3390/app11041697 - 14 Feb 2021
Cited by 6 | Viewed by 1575
Abstract
While the dependence assumption among the components is naturally important in evaluating the reliability of a system, studies investigating the issues of aggregation errors in Bayesian reliability analyses have been focused mainly on systems with independent components. This study developed a copula-based Bayesian [...] Read more.
While the dependence assumption among the components is naturally important in evaluating the reliability of a system, studies investigating the issues of aggregation errors in Bayesian reliability analyses have been focused mainly on systems with independent components. This study developed a copula-based Bayesian reliability model to formulate dependency between components of a parallel system and to estimate the failure rate of the system. In particular, we integrated Monte Carlo simulation and classification tree learning to identify key factors that affect the magnitude of errors in the estimation of posterior means of system reliability (for different Bayesian analysis approaches—aggregate analysis, disaggregate analysis, and simplified disaggregate analysis) to provide important guidelines for choosing the most appropriate approach for analyzing a model of products of a probability and a frequency for parallel systems with dependent components. Full article
(This article belongs to the Section Applied Industrial Technologies)
Show Figures

Figure 1

Figure 1
<p>Observed CDF of individual parameters in AA−DA analysis.</p>
Full article ">Figure 2
<p>Observed CDF of individual parameters in DAI−DA analysis.</p>
Full article ">Figure 3
<p>Classification tree obtained from adjusted parameters.</p>
Full article ">
22 pages, 3532 KiB  
Article
A Tail Dependence-Based MST and Their Topological Indicators in Modeling Systemic Risk in the European Insurance Sector
by Anna Denkowska and Stanisław Wanat
Risks 2020, 8(2), 39; https://doi.org/10.3390/risks8020039 - 22 Apr 2020
Cited by 7 | Viewed by 3603
Abstract
In the present work, we analyze the dynamics of indirect connections between insurance companies that result from market price channels. In our analysis, we assume that the stock quotations of insurance companies reflect market sentiments, which constitute a very important systemic risk factor. [...] Read more.
In the present work, we analyze the dynamics of indirect connections between insurance companies that result from market price channels. In our analysis, we assume that the stock quotations of insurance companies reflect market sentiments, which constitute a very important systemic risk factor. Interlinkages between insurers and their dynamics have a direct impact on systemic risk contagion in the insurance sector. Herein, we propose a new hybrid approach to the analysis of interlinkages dynamics based on combining the copula-DCC-GARCH model and minimum spanning trees (MST). Using the copula-DCC-GARCH model, we determine the tail dependence coefficients. Then, for each analyzed period we construct MST based on these coefficients. The dynamics are analyzed by means of the time series of selected topological indicators of the MSTs in the years 2005–2019. The contribution to systemic risk of each institution is determined by analyzing the deltaCoVaR time series using the copula-DCC-GARCH model. Our empirical results show the usefulness of the proposed approach to the analysis of systemic risk (SR) in the insurance sector. The times series obtained from the proposed hybrid approach reflect the phenomena occurring in the market. We check whether the analyzed MST topological indicators can be considered as systemic risk predictors. Full article
(This article belongs to the Special Issue Systemic Risk and Reinsurance)
Show Figures

Figure 1

Figure 1
<p>Degree distribution of selected minimum spanning trees. Source: Own study.</p>
Full article ">Figure 2
<p>Average value of betweenness centrality (BC) in the period under consideration for individual insurance institutions. Source: Own study.</p>
Full article ">Figure 3
<p>Average insurers strength in the period under consideration. Source: Own study.</p>
Full article ">Figure 4
<p>Average closeness centrality in the period under consideration. Source: Own study.</p>
Full article ">Figure 5
<p>Average betweenness centrality, vertex degree, vertex strength, and closeness centrality. Source: Own study.</p>
Full article ">Figure 6
<p>Average path length in the period from 7 January 2005 to 20 December 2019. Red lines depict the 13 periods of the moving average smoothed series. Source: Own study.</p>
Full article ">Figure 7
<p>Maximum degree in the period from 7 January 2005 to 20 December 2019. Source: Own study.</p>
Full article ">Figure 8
<p>Estimated parameters alpha of power distribution for the minimum spanning trees (MST) from 7 January 2005 to 20 December 2019. Source: Own study.</p>
Full article ">Figure 9
<p>Diameter in the period from 7 January 2005 to 20 December 2019. Red lines depict the 13 periods of the moving average smoothed series. Source: Own study.</p>
Full article ">Figure 10
<p>Rich club effect (RCE) during the period from 7 January 2005 to 20 December 2019. Red lines depict the 13 periods of the moving average smoothed series. Source: Own study.</p>
Full article ">Figure 11
<p>Assortativity during the period from 7 January 2005 to 20 December 2019. Red lines depict the 13 periods of the moving average smoothed series. Source: Own study.</p>
Full article ">Figure 12
<p>Distribution of MST topological indicators in different market states. Source: Own study.</p>
Full article ">Figure 13
<p>RCE distribution in determined market states based on the mean and the standard deviation for k = 4. Source: Own study.</p>
Full article ">Figure 14
<p>Sample MSTs in two chosen times. Source: Own study.</p>
Full article ">Figure 14 Cont.
<p>Sample MSTs in two chosen times. Source: Own study.</p>
Full article ">Figure 15
<p>Mean deltaCoVaR. Source: Own study.</p>
Full article ">Figure 16
<p>Mean deltaCoVaR for each insurer. Source: Own study.</p>
Full article ">Figure 17
<p>Mean deltaCoVaR distribution in different market states. Source: Own study.</p>
Full article ">Figure 17 Cont.
<p>Mean deltaCoVaR distribution in different market states. Source: Own study.</p>
Full article ">
22 pages, 1179 KiB  
Article
Direct and Hierarchical Models for Aggregating Spatially Dependent Catastrophe Risks
by Rafał Wójcik, Charlie Wusuo Liu and Jayanta Guin
Risks 2019, 7(2), 54; https://doi.org/10.3390/risks7020054 - 8 May 2019
Cited by 1 | Viewed by 7596
Abstract
We present several fast algorithms for computing the distribution of a sum of spatially dependent, discrete random variables to aggregate catastrophe risk. The algorithms are based on direct and hierarchical copula trees. Computing speed comes from the fact that loss aggregation at branching [...] Read more.
We present several fast algorithms for computing the distribution of a sum of spatially dependent, discrete random variables to aggregate catastrophe risk. The algorithms are based on direct and hierarchical copula trees. Computing speed comes from the fact that loss aggregation at branching nodes is based on combination of fast approximation to brute-force convolution, arithmetization (regriding) and linear complexity of the method for computing the distribution of comonotonic sum of risks. We discuss the impact of tree topology on the second-order moments and tail statistics of the resulting distribution of the total risk. We test the performance of the presented models by accumulating ground-up loss for 29,000 risks affected by hurricane peril. Full article
Show Figures

Figure 1

Figure 1
<p>Aggregation of five risks using copula trees. Direct model (upper panel), hierarchical model with sequential topology (middle panel) and hierarchical model with closest pair topology (lower panel). The leaf nodes represent the risks whose aggregate we are interested in. The branching nodes of direct tree represent a multivariate copula model for the incoming individual risks while the branching nodes of hierarchical trees represent a bivariate copula model for the incoming pairs of individual and/or cumulative risks.</p>
Full article ">Figure 2
<p>Illustration of hypothetical correlation matrices: (<b>A</b>) exchangeable, (<b>B</b>) nested block diagonal, and (<b>C</b>) unstructured correlation matrix.</p>
Full article ">Figure 3
<p>A discrete loss pdf represented as a mixture of two “spikes” (atoms) at minimum and maximum x damage ratio (red) and the main part (blue). Damage ratio is discretized on 64-point grid.</p>
Full article ">Figure 4
<p>An example of RNN approach for determining topology of hierarchical risk aggregation tree for six risks with zero minima. The maxima and cumulative maxima characterizing losses for the six risks are presented in the upper panel. (<b>A</b>) The algorithm takes the largest cumulative max and halves it to obtain the number <span class="html-italic">c</span>. Then, it binary searches for the number closest to <span class="html-italic">c</span> except for the last element in the sequence. This number (showed in bold) becomes the cumulative maximum of the new subsequence. The search is repeated until the subsequence consists of two elements. (<b>B</b>) The resulting hierarchical aggregation tree.</p>
Full article ">Figure 5
<p>An example of recursive nearest neighbor (RNN) approach for determining topology of direct risk aggregation tree for six risks shown in <a href="#risks-07-00054-f004" class="html-fig">Figure 4</a>. Note that the order of comonotonic aggregation follows the order of independent aggregation.</p>
Full article ">Figure 6
<p>MC (red line) and convolution/comonotoinization based (blue bars) distributions of the total risk for 29,139 locations affected by hurricane peril using different aggregation models with linear regriding (upper row) and 4-point regriding (lower row). No tail truncation was applied. For consistency, the losses are plotted in [0; <span>$</span>100 MM] interval.</p>
Full article ">Figure 7
<p>(<b>A</b>–<b>E</b>) Percentage errors in statistics of the total risk relative to the corresponding values obtained from MC simulations. Risk aggregation was performed for 29,139 locations affected by hurricane peril using sequential (blue) and RNN (red) models with 4-point regriding and maximum support size varying from 64 to 6400; (<b>F</b>) shows the average time cost of five runs for each maximum support size.</p>
Full article ">
17 pages, 2528 KiB  
Article
Application of Copula Functions for Rainfall Interception Modelling
by Nejc Bezak, Katarina Zabret and Mojca Šraj
Water 2018, 10(8), 995; https://doi.org/10.3390/w10080995 - 27 Jul 2018
Cited by 13 | Viewed by 3728
Abstract
Rainfall interception is an important process of the water cycle that can have significant influence on surface runoff and groundwater storage. Since rainfall interception measurements are rare and time consuming, rainfall interception estimation can be made indirectly using different meteorological variables. Experimental data [...] Read more.
Rainfall interception is an important process of the water cycle that can have significant influence on surface runoff and groundwater storage. Since rainfall interception measurements are rare and time consuming, rainfall interception estimation can be made indirectly using different meteorological variables. Experimental data of rainfall interception for birch and pine trees was measured at an experimental plot located in an urban area of Ljubljana, Slovenia in this study. A copula model was applied to predict the rainfall interception using meteorological variables, namely air temperature and vapour pressure deficit data. The copula model performance was compared to some other models such as decision trees, multiple linear regressions, and exponential functions. Using random sampling, we found that the copula model where Khoudraji-Liebscher copula functions were used yielded slightly smaller root mean square error (RMSE) and mean absolute error (MAE) values than other tested methods (i.e., RMSE and MAE results for birch trees were 24.2% and 18.2%, respectively and RMSE and MAE results for pine trees were 25.0% and 19.6%, respectively). The results demonstrate that the copula-based proposed method and other tested models could be used for the prediction of rainfall interception at the considered plot and in the wider surroundings. Furthermore, these models could also be applied for the prediction of rainfall interception for these two tree species in other locations under similar vegetation and meteorological conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Slovenia on the map of Europe, and a photo from the measuring plot showing the measuring equipment used.</p>
Full article ">Figure 2
<p>Graphical presentation of the relationship among interception loss for pine trees (<span class="html-italic">Ip</span>), vapour pressure deficit and air temperature (<span class="html-italic">T</span>) (upper panel) and interception loss for birch trees (<span class="html-italic">Ib</span>), vapour pressure deficit and air temperature (<span class="html-italic">T</span>) (lower panel). Events measured in the leafless and leafed period are indicated with red and black colour, respectively.</p>
Full article ">Figure 3
<p>Rainfall interception by birch (<span class="html-italic">Ib</span>) and pine (<span class="html-italic">Ip</span>) trees according to the amount of rainfall in the open.</p>
Full article ">Figure 4
<p>Graphical fit for birch trees between measured data (red crosses) and 10,000 random realizations (grey circles) using the copula model where the Joe copula was selected as <span class="html-italic">C</span><sub>1</sub> and the Gumbel-Hougaard copula was selected as <span class="html-italic">C</span><sub>2</sub>.</p>
Full article ">Figure 5
<p>Graphical fit for pine trees between measured data (red crosses) and 10,000 random realizations (grey circles) using the copula model where the Joe copula was selected as C<sub>1</sub> and the Gumbel-Hougaard copula was selected as C<sub>2</sub>.</p>
Full article ">Figure 6
<p>Decision trees for rainfall interception by birch (<b>a</b>) and pine (<b>b</b>) trees; only the first three levels are presented. Values shown in boxes indicate mean and variance of instances related to this specific node. Numbers written outside boxes indicate values that are used to divide different nodes.</p>
Full article ">
Back to TopTop