\chapter{Fundamentals of Finite Element Method for structural applications} \section{Introduction} \todo{Missing.}Briefly.\\ %\begin{figure} % \centering % \includegraphics{prova_multilingue/fig.pdf} % \label{fig} % \caption{hello} %\end{figure} \section{Preparatory formulations} \subsection{Interpolation functions for some elementary domains} \paragraph{Quadrilateral domain.} A quadrilateral domain is considered whose vertices are conventionally located at the $[-1,-1]$, $[1,-1]$, $[1,1]$ and $[-1,1]$ points of an adimensional $[\xi,\eta]$ plane coordinate system, see Figure. \begin{figure} \label{fig:iso4shapefun} \centering \includegraphics{images/iso4_shapefun.pdf} \caption{Quadrilateral elementary domain (a), and a representative weight function (b).} \end{figure} Scalar values $f_i$ are associated to a set of \emph{nodal} points $\pt{P}_i\equiv[\xi_i,\eta_i]$, which for the present case coincide with the quadrangle vertices, numbered as in figure. A $f(\xi,\eta)$ interpolation function may be devised by defining a set of nodal influence functions $N_i(\xi,\eta)$ to be employed as the coefficients (weights) of a moving weighted average \begin{equation} f(\xi,\eta) \defeq \sum_i N_i(\xi,\eta) f_i \label{interpfun} \end{equation} Requisite for such weight functions are that \begin{itemize} \item the influence of a node is unitary at its location, whereas other nodes influence locally vanishes \begin{equation} N_i(\xi_j,\eta_j)=\delta_{ij} \label{weightfunatnodes} \end{equation} \item for each point of the domain, the sum of the weights is unitary \begin{equation} \sum_i N_i(\xi,\eta) = 1, \; \forall [\xi,\eta] \label{weighteddsum} \end{equation} \end{itemize} Moreover, functions should be continuous and straightforwardly differentiable up to any required degree. Low order polynomials are ideal candidates for the application; for the particular element, the nodal weight functions may be stated as \begin{equation} N_i(\xi,\eta) \defeq \frac{1}{4} \left(1 \pm \xi \right) \left(1 \pm \eta \right), \label{quad4InterpFunDef} \end{equation} where sign ambiguity is resolved for each $i$-th node by enforcing Eqn. \ref{weightfunatnodes}. The (\ref{weighteddsum}) combination of \ref{quad4InterpFunDef} turns into a generic linear relation in $(\xi,\eta)$ with coplanar -- but otherwise arbitrary -- $(\xi_i,\eta_i,f_i)$ points. Further generality may be introduced by \emph{not} enforcing coplanarity. The weight functions for the four-node quadrilateral are in fact quadratically incomplete\footnote{or \emph{enriched linear}, as discussed above and in the following} in nature due to the presence of the $\xi\eta$ product, and the absence of any $\xi^2,\eta^2$ term. Each term, and the combined interpolation function $f(\xi,\eta)$, defined as in Eqn. \ref{interpfun}, behave linearly if restricted to $\xi=\mathrm{const.}$ or $\eta=\mathrm{const.}$ loci -- namely along the four edges, whereas quadratic behaviour may arise along a general direction, e.g. along the diagonals, as in the Fig. \ref{fig:iso4shapefun}b example. Such behaviour is called \emph{bilinear}. We now consider the $f(\xi,\eta)$ weight function partial derivatives. The partial derivative \begin{equation} \frac{\partial f}{\partial \xi} = \underbrace{ \left(\frac{f_2-f_1}{2}\right) }_{\left[\Delta f / \Delta \xi\right]_{12}} \underbrace{ \left(\frac{1-\eta}{2} \right) }_{N_1+N_2} + \underbrace{ \left(\frac{f_3-f_4}{2}\right) }_{\left[\Delta f / \Delta \xi\right]_{43}} \underbrace{ \left(\frac{1+\eta}{2}\right) }_{N_4+N_3} = a \eta + b \end{equation} linearly varies from the incremental ratio value measured at the $\eta=-1$ lower edge, to the value measured at the $\eta=1$ upper edge; the other partial derivative \begin{equation} \frac{\partial f}{\partial \eta} = \left(\frac{f_4-f_1}{2}\right) \left(\frac{1-\xi}{2}\right) + \left(\frac{f_3-f_2}{2}\right) \left(\frac{1+\xi}{2}\right) = c \xi + d . \end{equation} similarly behaves, with $c=a$. However, partial derivatives in $\xi,\eta$ remain constant along the corresponding differentiation direction \footnote{The relevance of such partial derivative orders will appear clearer to the reader once the strain field will have been derived in paragraph XXX.}. \subsection{Gaussian quadrature rules for reference domains.} \paragraph{Reference one dimensional domain.} The gaussian quadrature rule for approximating the definite integral of a $f(\xi)$ function over the $[-1,1]$ reference interval is constructed as the customary weighted sum of internal function samples, namely \begin{equation} \int_{-1}^{1} f(\xi) d \xi \approx \sum_{i=1}^n f(\xi_i) w_i; \label{gaussQuadDef} \end{equation} Its peculiarity is to employ location-weight pairs $\left(\xi_i,w_i\right)$ that are optimal with respect to the polynomial class of functions. Nevertheless, such choice has revealed itself to be robust enough for a more general use. Let's consider an $m$-th order polynomial \begin{equation*} p(\xi) \defeq a_m \xi^m + a_{m-1} \xi^{m-1} + \ldots + a_1 \xi +a_0 \end{equation*} whose exact integral is \begin{equation*} \int_{-1}^{1} p(\xi) d \xi = \sum_{j=0}^m \frac{(-1)^{j}+1}{j+1} a_j \end{equation*} The integration residual between the exact definite integral and the weighted sample sum is defined as \begin{equation} r \left(a_j, \left(\xi_i,w_i \right) \right) \defeq \sum_{i=1}^n p(\xi_i) w_i - \int_{-1}^{1} p(\xi) d \xi \label{quadResidual} \end{equation} The optimality condition is stated as follows: the quadrature rule involving $n$ sample points $\left(\xi_i,w_i \right)$, $i=1 \ldots n$ is optimal for the $m$-th order polynomial if a) the integration residual is null for general $a_j$ values , and b) such condition does not hold for any lower-order sampling rule. Once observed that the zero residual requirement is satisfied by any sampling rule if the polynomial $a_j$ coefficients are all null, condition a) may be enforced by imposing that such zero residual value remains constant with varying $a_j$ terms, i.e. \begin{equation} \left\lbrace\frac{\partial r \left(a_j, \left(\xi_i,w_i \right) \right)}{\partial a_j}=0, \quad j=0\ldots m \right. \label{statQuadResidual} \end{equation} A system of $m+1$ polynomial equations of degree $m-1$ is hence obtained in the $2n$ $(\xi_i,w_i)$ unknowns; in the assumed absence of redundant equations, solutions are not allowed if the constraints outnumber the unknowns, i.e. $m > 2n - 1$. Limiting our discussion to the threshold condition $m = 2n - 1$, an attentive algebraic manipulation of Eqns. \ref{statQuadResidual} may be performed in order to extract the $(\xi_i,w_i)$ solutions, which are unique apart from the pair permutations\footnote{ In this note, location-weight pairs are obtained for the gaussian quadrature rule of order $n=2$, aiming at illustrating the general procedure. The general $m=2n-1=3$rd order polynomial is stated in the form $$ p(\xi) = a_3 \xi^3 + a_2 \xi^2 + a_1 \xi +a_0, \quad \int_{-1}^{1} p(\xi) d \xi = \frac{2}{3} a_2 + 2 a_0, $$ whereas the integral residual is $$ r = a_3 \left(w_1\xi_1^3+w_2\xi_2^3 \right) + a_2 \left(w_1\xi_1^2+w_2\xi_2^2-\frac{2}{3} \right) + a_1 \left(w_1\xi_1 +w_2\xi_2 \right) + a_0 \left(w_1 +w_2 -2 \right) $$ Eqns \ref{statQuadResidual} may be derived as $$ \left\lbrace \begin{array}{ll} 0=\frac{\partial r}{\partial a_3} = w_1\xi_1^3+w_2\xi_2^3 & (e_1)\\ 0=\frac{\partial r}{\partial a_2} = w_1\xi_1^2+w_2\xi_2^2-\frac{2}{3} & (e_2)\\ 0=\frac{\partial r}{\partial a_1} = w_1\xi_1 +w_2\xi_2 & (e_3)\\ 0=\frac{\partial r}{\partial a_0} = w_1 +w_2 -2 & (e_4) \end{array} \right. $$ which are independent of the $a_j$ coefficients. By composing $\left(e_1 - \xi_1^2 e_3\right)/(w_2 \xi_2)$ it is obtained that $\xi_2^2 = \xi_1^2$; $e_2$ may then be written as $(w_1+w_2)\xi_1^2=2/3$, and then as $2\xi_1^2=2/3$, according to $e_4$. It derives that $\xi_{1,2}=\pm 1/\sqrt{3}$. Due to the opposite nature of the roots, $e_3$ implies $w_2=w_1$, relation, this, that turns $e_4$ into $2w_1=2w_2=2$, and hence $w_{1,2}=1$ . }. \begin{table} \centering \begin{tabu}{ l @{\hspace{2em}} l @{\hspace{2em}} l } \toprule $n$ & $\xi_i$ & $w_i$ \\ \midrule 1 & 0 & 2 \\ \midrule 2 & $\pm \frac{1}{\sqrt{3}} $ & 1 \\ \midrule \multirow{2}{*}{3} & 0 & $\frac{8}{9}$ \\ %\cmidrule{2-3} & $\pm \sqrt{\frac{3}{5}} $ & $\frac{5}{9}$ \\ \midrule \multirow{2}{*}{4} & $\pm \sqrt{\frac{3}{7}-\frac{2}{7}\sqrt{\frac{6}{5}}} $ & $\frac{18+\sqrt{30}}{36}$ \\ %\cmidrule{2-3} & $\pm \sqrt{\frac{3}{7}+\frac{2}{7}\sqrt{\frac{6}{5}}} $ & $\frac{18-\sqrt{30}}{36}$ \\ \bottomrule \end{tabu} \caption{Integration points for the lower order gaussian quadrature rules.} \label{table:gaussPoints} \end{table} Eqns. \ref{statQuadResidual} solutions are reported in Table \ref{table:gaussPoints} for quadrature rules with up to $n=4$ sample points\footnote{see \url{https://pomax.github.io/bezierinfo/legendre-gauss.html} for higher order gaussian quadrature rule sample points.}. It is noted that the integration points are symmetrically distributed with respect to the origin, and that the function is never sampled at the $\left\lbrace-1,1\right\rbrace$ extremal points. \paragraph{General one dimensional domain.} The extension of the one dimensional quadrature rule from the reference domain $[-1,1]$ to a general $[a,b]$ domain is pretty straightforward, requiring just a change of integration variable to obtain the following \begin{align*} \int_{a}^{b} f(x) d x &= \frac{b-a}{2} \int_{-1}^{1} f \left( \frac{b+a}{2} + \frac{b-a}{2}\xi \right) d\xi, \\ &\approx \frac{b-a}{2} \sum_{i=1}^{n} f \left( \frac{b+a}{2} + \frac{b-a}{2}\xi_i \right) w_i. \end{align*} \paragraph{Reference quadrangular domain.} A quadrature rule for the reference quadrangular domain of Figure XXX may be derived by nesting the quadrature rule defined for the reference interval, see Eqn. \ref{gaussQuadDef}, thus obtaining \begin{equation} \int_{-1}^{1} \int_{-1}^{1} f \left( \xi,\eta \right) d \xi d \eta \approx \sum_{i=1}^{p}\sum_{j=1}^{q} f \left( \xi_i,\eta_j \right) w_i w_j \end{equation} where $(\xi_i , w_i)$ and $(\xi_j , w_j)$ are the coordinate-weight pairs of the two quadrature rules of $p$ and $q$ order employed for spanning the two coordinate axes. The equivalent notation \begin{equation} \int_{-1}^{1} \int_{-1}^{1} f \left( \xi,\eta \right) d \xi d \eta \approx \sum_{l=1}^{pq} f \left( \vct{\xi}_l \right) w_l \label{eq:gaussQuadDef_quad} \end{equation} emphasises the characteristic nature of the $pq$ point/weight pairs for the domain and for the quadrature rule employed; a general integer biiection $\lbrace 1 \ldots pq \rbrace \leftrightarrow \lbrace 1 \ldots p \rbrace \times \lbrace 1 \ldots q \rbrace$, $ l \leftrightarrow (i,j)$ e.g. \begin{equation} \lbrace i -1 ; j - 1 \rbrace = ( l-1 ) \mmod ( p , q ) , \quad l-1 = (j-1) q + (i-1) \end{equation} may be utilized to formally derive the two-dimensional quadrature rule pairs \begin{equation} \vct{\xi}_l = \left( \xi_i , \eta_j \right) , \quad w_l = w_i w_j ,\quad l=1\ldots pq \end{equation} from their uniaxial counterparts. \paragraph{General quadrangular domain.} The interpolation functions introduced in Paragraph \ref{quad4InterpFun} may be be profitably employed in defining a coordinate mapping between a general quadrangular domain -- see Fig. XXXx -- and its reference counterpart. In particular, we first define the $\vct{\xi}_i\mapsto\vct{x}_i$ mapping for the coordinates of the four vertices\footnote{The condensed notation $\vct{\xi}_i\equiv \left(\xi_i,\eta_i\right)$, $\vct{x}_i\equiv\left(x_i,y_i\right)$ for coordinate vectors is employed.} in the natural (or reference) $\xi,\eta$ and in the physical $x,y$ reference systems. Then, a mapping for the inner points may be derived by interpolation, namely \begin{equation} \vct{x}\left( \vct{\xi} \right) = \sum_{i=1}^{4} N_i \left( \vct{\xi}\right) \vct{x}_i \end{equation} The availability of an inverse $\vct{\xi}\left( \vct{x} \right)$ mapping is not granted; in particular, a closed form representation for such inverse is not generally available\footnote{For a given $\bar{x}$ physical point, however, Newton-Raphson iterations rapidly converge to the $\vct{\xi}\left(\vct{\bar{x}}\right)$ solution if the centroid is supplied for algorithm initialization, see Section XXX}. The rectangular infinitesimal area $dA_{\xi\eta}=d\xi d\eta$ in the neighborhood of a $\xi^\star,\eta^\star$ location, see Figure XXXa, is mapped to the quadrangle of Figure XXXb, which is composed by the two triangular areas { \newcommand{\nulldxi} {\mathbin{\phantom{+}}\phantom{d\xi}} \newcommand{\nulldeta}{\mathbin{\phantom{+}}\phantom{d\eta}} \begin{align} dA_{xy} = &\mathbin{\phantom{+}} \frac{1}{2!} \begin{vmatrix} 1 & x\left( \xi^\star\nulldxi , \eta^\star\nulldeta \right) & y\left( \xi^\star\nulldxi , \eta^\star\nulldeta \right) \\ 1 & x\left( \xi^\star+d\xi , \eta^\star\nulldeta \right) & y\left( \xi^\star+d\xi , \eta^\star\nulldeta \right) \\ 1 & x\left( \xi^\star\nulldxi , \eta^\star+d\eta \right) & y\left( \xi^\star\nulldxi , \eta^\star+d\eta \right) \end{vmatrix} + \nonumber\\ &+ \frac{1}{2!} \begin{vmatrix} 1 & x\left( \xi^\star+d\xi , \eta^\star+d\eta \right) & y\left( \xi^\star+d\xi , \eta^\star+d\eta \right) \\ 1 & x\left( \xi^\star\nulldxi , \eta^\star+d\eta \right) & y\left( \xi^\star\nulldxi , \eta^\star+d\eta \right) \\ 1 & x\left( \xi^\star+d\xi , \eta^\star\nulldeta \right) & y\left( \xi^\star+d\xi , \eta^\star\nulldeta \right) \end{vmatrix} \label{eq:dAxy} \end{align} } where the determinant formula for the area of a triangle, shown below along with its $n$-dimensional symplex hypervolume generalization \begin{equation} \mathcal{A} = \frac{1}{2!} \begin{vmatrix} 1 & x_1 & y_1 \\ 1 & x_2 & y_2 \\ 1 & x_3 & y_3 \end{vmatrix} ,\quad \mathcal{H} = \frac{1}{n!} \begin{vmatrix} 1 & \vct{x}_1 & \\ 1 & \vct{x}_2 & \\ \vdots & \vdots \\ 1 & \vct{x}_{n+1} \\ \end{vmatrix} \end{equation} is employed. By operating a local multivariate linearization of \ref{eq:dAxy} matrix terms, \begin{align*} dA_{xy} \approx &\mathbin{\phantom{+}} \frac{1}{2!} \begin{vmatrix} 1& x^\star & y^\star \\ 1& x^\star + x^\star_{,\xi} d\xi & y^\star + y^\star_{,\xi} d\xi \\ 1& x^\star + x^\star_{,\eta} d\eta& y^\star + y^\star_{,\eta} d\eta \end{vmatrix} + \nonumber\\ &+ \frac{1}{2!} \begin{vmatrix} 1& x^\star + x^\star_{,\xi} d\xi + x^\star_{,\eta} d\eta& y^\star + y^\star_{,\xi} d\xi + y^\star_{,\eta} d\eta \\ 1& x^\star + x^\star_{,\eta} d\eta& y^\star + y^\star_{,\eta} d\eta \\ 1& x^\star + x^\star_{,\xi} d\xi & y^\star + y^\star_{,\xi} d\xi \end{vmatrix} \label{eq:dAxyLin} \end{align*} is obtained, where $x^\star,y^\star$ and $x^\star_{,\xi},x^\star_{,\eta},y^\star_{,\xi},y^\star_{,\eta}$ are the $x,y$ functions and their first order partial derivatives, as sampled at the $\xi^\star,\eta^\star$ point; infinitesimal terms of higher order than $d\xi d\eta$ are neglected . After some matrix manipulations\footnote{ For both the determinants, the first column is multiplied by $x^\star$ and subtracted to the second column, and then subtracted to the third column once multiplied by $y^\star$. The first row is then subtracted to the others. On the second determinant alone, both the second and the third columns are changed in sign; then, the second and the third rows are summed to the first. The two determinants are now formally equal, and the two $1/2$ multipliers are summed to unity. The $d\xi$ and the $d\eta$ factors may then be collected from the second and the third rows, respectively. } the following expression is obtained \begin{equation} dA_{xy} = \begin{vmatrix} 1& 0 & 0 \\ 0& x^\star_{,\xi}& y^\star_{,\xi}\\ 0& x^\star_{,\eta}& y^\star_{,\eta} \end{vmatrix} d\xi d\eta = \underbrace{ \begin{vmatrix} x^\star_{,\xi}& y^\star_{,\xi} \\ x^\star_{,\eta}& y^\star_{,\eta} \end{vmatrix} }_{\left| J^\mathrm{T}(\xi,\eta) \right|} dA_{\xi\eta} \end{equation} that equates the ratio of the mapped and the original areas to the determinant of the transformation (transpose) Jacobian matrix\footnote{The Jacobian matrix for a general $\vct{\xi} \mapsto \vct{x}$ mapping is in fact defined according to \[ [J( \vct{\xi}^\star )]_{ij} \defeq \left. \frac{\partial x_i}{\partial \xi_j} \right|_{\vct{\xi}=\vct{\xi}^\star} \quad i,j = 1 \ldots n \] being $i$ the generic matrix term row index, and $j$ the column index}. Once carried out the preparatory passages, we obtain \begin{equation} \iint_{A_{xy}} g(x,y) dA_{xy} = \iint_{-1}^{1} g \left( x \left(\xi,\eta\right), y \left(\xi,\eta\right) \right) \left| J(\xi,\eta) \right| d\xi d\eta, \end{equation} thus reducing the quadrature over a general domain to its reference domain counterpart, which has been discussed in the paragraph above. Based on Eqn. \ref{eq:gaussQuadDef_quad}, the quadrature rule \begin{equation} \iint_{A_{xy}} g(\vct{x}) dA_{xy} \approx \sum_{l=1}^{pq} g \left( \vct{x} \left( \vct{\xi}_l \right) \right) \left| J(\vct{\xi}_l) \right| w_l \end{equation} is derived, stating that the definite integral of a $g$ integrand over a quadrangular domain pertaining to a physical $x,y$ plane ($x,y$ may be dimensional quantities, namely lengths) may be approximated by \begin{enumerate} \item defining a reference to physical domain mapping based on vertex physical coordinates interpolation; \item sampling the function at physical locations which are the images of the Gaussian integration points obtained for the reference domain; \item proceeding with a weighted sum of the collected samples, where the weights consist in the product of the adimensional $w_l$ Gauss point weight (suitable for integrating on the reference domain), and of a dimensional area scaling term consisting in the determinant of the transformation Jacobian matrix, as evaluated locally at the Gauss point. \end{enumerate}