Comparison of programming paradigms: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>BattyBot
 
(One intermediate revision by one other user not shown)
Line 1: Line 1:
{{lead rewrite|date=January 2010}}
{{Continuum mechanics|cTopic=[[Solid mechanics]]}}
In [[continuum mechanics]], a '''compatible''' [[finite deformation tensor|deformation]] (or [[strain tensor|strain]]) tensor field in a body is that ''unique'' field that is obtained when the body is subjected to a [[continuous function|continuous]], [[single-valued]], [[displacement field (mechanics)|displacement field]].  '''Compatibility''' is the study of the conditions under which such a displacement field can be guaranteed.  Compatibility conditions are particular cases of [[integrability condition]]s and were first derived for [[linear elasticity]] by [[Adhémar Jean Claude Barré de Saint-Venant|Barré de Saint-Venant]] in 1864 and proved rigorously by [[Eugenio Beltrami|Beltrami]] in 1886.<ref name="Amrouche">C Amrouche, [[Philippe G. Ciarlet|PG Ciarlet]], L Gratie, S Kesavan,  On Saint Venant's compatibility conditions and Poincaré's lemma, C. R. Acad. Sci. Paris, Ser. I, 342 (2006), 887-891.  {{doi|10.1016/j.crma.2006.03.026}}</ref>


In the continuum description of a solid body we imagine the body to be composed of a set of infinitesimal volumes or material points.  Each volume is assumed to be connected to its neighbors without any gaps or overlaps.  Certain mathematical conditions have to be satisfied to ensure that gaps/overlaps do not develop when a continuum body is deformed.  A body that deforms without developing any gaps/overlaps is called a '''compatible''' body.  '''Compatibility conditions''' are mathematical conditions that determine whether a particular deformation will leave a body in a compatible state.<ref name=Barber>Barber, J. R., 2002, Elasticity - 2nd Ed., Kluwer Academic Publications.</ref> 


In the context of [[infinitesimal strain theory]], these conditions are equivalent to stating that the displacements in a body can be obtained by integrating the [[infinitesimal strain theory|strain]]s. Such an integration is possible if the Saint-Venant's tensor (or incompatibility tensor) <math>\boldsymbol{R}(\boldsymbol{\varepsilon})</math> vanishes in a [[simply connected|simply-connected body]]<ref>N.I. Muskhelishvili,  Some Basic Problems of the Mathematical Theory of Elasticity. Leyden: Noordhoff Intern. Publ., 1975.</ref> where <math>\boldsymbol{\varepsilon}</math> is the [[infinitesimal strain theory|infinitesimal strain tensor]] and  
There are many types of bike brakes for different applications. Various means of transportation come second - from yachts or boats and cars to motorcycles. The contemporary bikes for mountains are provided with stronger and lighter frame types in addition to pioneering form and design. Mountain bike frames spend a lot of time on roads these days, too. If you've ridden for years and years and know what you need, buying online does make sense. <br><br>Once you are sure that this is what you want to do, it is now time for you to decide on the bike. Once you're comfortable coasting, dropping, standing, pedaling, spinning, and switching gears, you'll be ready to hit the trails, and tackle any challenge along the way. So how exactly would you go about making this dream a reality. If biking is your passion or your favorite sport, then you know the worth of good accessories while riding. If you are riding through a rock garden, for instance, you don't want a fork that is bouncing you all over the place. <br><br>Some also use the liner as a size aid, for example BMX helmets can come as one-size, with optional liner kits to change the helmets size. An Indian travois style trailer is very simple to make and very effective for heavy loads. The whole set-up, including your bag, should weigh no more than 20kgs (44lbs). Another way to classify brakes is by mounting style. The most common sized crankset is the 170 millimeter. <br><br>There are still other types of bikes that require different types of tires. This article discusses how to go about buying a mountain bike. Just remember your heart is a muscle and can be developed the more you practice. With a Shimano 105 groupset and FSA Gossamer chain set, it's a great choice for under. I've never been into cycling and wanted to give it a try for exercise. <br><br>If you have any kind of concerns relating to where and the best ways to utilize [http://dencle.lonnieheart.gethompy.com/?document_srl=219948 Choosing the right ride for you mountain bike sizing.], you could call us at our own site. I ran with Hale and we also tried this method, it was effective. Instability: You are much more likely to lose your balance while on a big bike. There are good bikes available at reasonable prices, and there are some overpriced stinkers. Suspension: Road bikes are built with a sole purpose of providing greater speed; they do not possess this feature, although they have certain materials which absorb the shocks of the uneven roads. They will provide you with a honest price, yet be ready to pay in between $650 and $4,000 on dual suspensions and in between $470 and $670 for hardtails.
:<math>
  \boldsymbol{R} := \boldsymbol{\nabla}\times(\boldsymbol{\nabla}\times\boldsymbol{\varepsilon}) ~.
</math>
For [[finite strain theory|finite deformations]] the compatibility conditions take the form
:<math>
  \boldsymbol{R} :=\boldsymbol{\nabla}\times\boldsymbol{F} = \boldsymbol{0}
</math>
where <math>\boldsymbol{F}</math> is the [[deformation gradient]].
 
== Compatibility conditions for infinitesimal strains ==
The compatibility conditions in [[linear elasticity]] are obtained by observing that there are six strain-displacement relations that are functions of only three unknown displacements.  This suggests that the three displacements may be removed from the system of equations without loss of information.  The resulting expressions in terms of only the strains provide constraints on the possible forms of a strain field.
 
=== 2-dimensions ===
For two-dimensional, [[plane strain]] problems the strain-displacement relations are
:<math>
  \varepsilon_{11} = \cfrac{\partial u_1}{\partial x_1} ~;~~
  \varepsilon_{12} = \cfrac{1}{2}\left[\cfrac{\partial u_{1}}{\partial x_2} + \cfrac{\partial u_{2}}{\partial x_1}\right]~;~~
  \varepsilon_{22} = \cfrac{\partial u_{2}}{\partial x_2}
</math>
 
Combining these relations gives us the two-dimensional compatibility condition for strains
:<math>
  \cfrac{\partial^2 \varepsilon_{11}}{\partial x_2^2}
  - 2\cfrac{\partial^2 \varepsilon_{12}}{\partial x_1 \partial x_2}
  + \cfrac{\partial^2 \varepsilon_{22}}{\partial x_1^2} = 0
</math>
 
The only displacement field that is allowed by a compatible plane strain field is a '''plane displacement''' field, i.e., <math> \mathbf{u} = \mathbf{u}(x_1, x_2) </math>.
 
=== 3-dimensions ===
In three dimensions, in addition to two more equations of the form seen for two dimensions, there are
three more equations of the form
:<math>
  \cfrac{\partial^2 \varepsilon_{33}}{\partial x_1 \partial x_2} = \cfrac{\partial}{\partial x_3}\left[
  \cfrac{\partial \varepsilon_{23}}{\partial x_1} + \cfrac{\partial \varepsilon_{31}}{\partial x_2} -
  \cfrac{\partial \varepsilon_{12}}{\partial x_3}\right]
</math>
Therefore there are '''six''' different compatibility conditions.  We can write these conditions in index notation as<ref name=Slaughter>Slaughter, W. S., 2003, ''The linearized theory of elasticity'', Birkhauser</ref>
:<math>
  e_{ikr}~e_{jls}~\varepsilon_{ij,kl} = 0
</math>
where <math>e_{ijk}</math> is the [[permutation symbol]]. In direct tensor notation
:<math>
  \boldsymbol{\nabla}\times(\boldsymbol{\nabla}\times\boldsymbol{\varepsilon}) = \boldsymbol{0}
</math>
where the curl operator can be expressed in an orthonormal coordinate system as <math>\boldsymbol{\nabla}\times\boldsymbol{\varepsilon} = e_{ijk}\varepsilon_{rj,i}\mathbf{e}_k\otimes\mathbf{e}_r </math>. 
 
The second-order tensor
:<math>
  \boldsymbol{R} := \boldsymbol{\nabla}\times(\boldsymbol{\nabla}\times\boldsymbol{\varepsilon}) ~;~~ R_{rs} := e_{ikr}~e_{jls}~\varepsilon_{ij,kl}
</math>
is known as the '''incompatibility tensor'''.
 
== Compatibility conditions for finite strains ==
For solids in which the deformations are not required to be small, the compatibility conditions take the form
:<math>
  \boldsymbol{\nabla}\times\boldsymbol{F} = \boldsymbol{0}
</math>
where <math>\boldsymbol{F}</math> is the [[deformation gradient]].  In terms of components with respect to a Cartesian coordinate system we can write these compatibility relations as
:<math>
  e_{ABC}~\cfrac{\partial F_{iB}}{\partial X_A} = 0
</math>
This condition is '''necessary''' if the deformation is to be continuous and derived from the mapping <math>\mathbf{x} = \boldsymbol{\chi}(\mathbf{X},t)</math> (see [[Finite strain theory]]).  The same condition is also '''sufficient''' to ensure compatibility in a '''[[simply connected]]''' body.
 
=== Compatibility condition for the right Cauchy-Green deformation tensor ===
The compatibility condition for the [[finite strain theory|right Cauchy-Green deformation tensor]] can be expressed as
:<math>
  R^\gamma_{\alpha\beta\rho} :=
  \frac{\partial }{\partial X^\rho}[\Gamma^\gamma_{\alpha\beta}] -
  \frac{\partial }{\partial X^\beta}[\Gamma^\gamma_{\alpha\rho}] +
  \Gamma^\gamma_{\mu\rho}~\Gamma^\mu_{\alpha\beta} -
  \Gamma^\gamma_{\mu\beta}~\Gamma^\mu_{\alpha\rho} = 0
</math>
where <math>\Gamma^k_{ij}</math> is the [[curvilinear coordinates|Christoffel symbol of the second kind]].  The quantity <math>R^m_{ijk}</math> represents the mixed components of the [[Riemann-Christoffel curvature tensor]].
 
== The general compatibility problem ==
The problem of compatibility in continuum mechanics involves the determination of allowable single-valued continuous fields on simply connected bodies.  More precisely, the problem may be stated in the following manner.<ref name=acharya>Acharya, A., 1999, '' On Compatibility Conditions for the Left Cauchy–Green Deformation Field in Three Dimensions'', Journal of Elasticity, Volume 56, Number 2 , 95-105</ref>
 
[[Image:Displacement of a continuum.svg|400px|right|thumb|Figure 1. Motion of a continuum body.]]
Consider the deformation of a body shown in Figure 1.  If we express all vectors in terms of the reference coordinate system <math>\{(\mathbf{E}_1, \mathbf{E}_2, \mathbf{E}_3), O\}</math>, the displacement of a point in the body is given by
:<math>
  \mathbf{u} = \mathbf{x} - \mathbf{X} ~;~~ u_i = x_i - X_i
</math>
Also
:<math>
  \boldsymbol{\nabla} \mathbf{u} = \frac{\partial \mathbf{u}}{\partial \mathbf{X}} ~;~~
  \boldsymbol{\nabla} \mathbf{x} = \frac{\partial \mathbf{x}}{\partial \mathbf{X}}
</math>
 
What conditions on a given second-order tensor field <math>\boldsymbol{A}(\mathbf{X})</math> on a body are necessary and sufficient so that there exists a unique vector field <math>\mathbf{v}(\mathbf{X})</math> that satisfies
:<math>
  \boldsymbol{\nabla} \mathbf{v} = \boldsymbol{A} \quad \equiv \quad v_{i,j} = A_{ij}
</math>
 
===Necessary conditions===
For the necessary conditions we assume that the field <math>\mathbf{v}</math> exists and satisfies
<math>v_{i,j} = A_{ij}</math>. Then
:<math>
  v_{i,jk} = A_{ij,k} ~;~~ v_{i,kj} = A_{ik,j}
</math>
Since changing the order of differentiation does not affect the result we have
:<math>
  v_{i,jk} = v_{i,kj}
</math>
Hence
:<math>
  A_{ij,k} = A_{ik,j}
</math>
From the well known identity for the [[tensor derivative (continuum mechanics)|curl of a tensor]] we get the necessary condition
:<math>
  \boldsymbol{\nabla} \times \boldsymbol{A} = \boldsymbol{0}
</math>
 
===Sufficient conditions===
[[Image:Compatibility mechanics.png|200px|right|thumb|Figure 2. Integration paths used in proving the sufficiency conditions for compatibility.]]
To prove that this condition is sufficient to guarantee existence of a compatible second-order tensor field, we start with the assumption that a field <math>\boldsymbol{A}</math> exists such that
<math> \boldsymbol{\nabla} \times \boldsymbol{A} = \boldsymbol{0} </math>. We will integrate this field to find the vector field <math>\mathbf{v}</math> along a line between points <math>A</math> and <math>B</math> (see Figure 2), i.e.,
:<math>
  \mathbf{v}(\mathbf{X}_B) - \mathbf{v}(\mathbf{X}_A) = \int_{\mathbf{X}_A}^{\mathbf{X}_B} \boldsymbol{\nabla} \mathbf{v}\cdot~d\mathbf{X}
  = \int_{\mathbf{X}_A}^{\mathbf{X}_B} \boldsymbol{A}(\mathbf{X})\cdot d\mathbf{X}
</math>
If the vector field <math>\mathbf{v}</math> is to be single-valued then the value of the integral should be independent of the path taken to go from <math>A</math> to <math>B</math>.
 
From [[Stokes theorem]], the integral of a second order tensor along a closed path is given by
:<math>
  \oint_{\partial\Omega} \boldsymbol{A}~ds = \int_{\Omega} \mathbf{n}\cdot(\boldsymbol{\nabla} \times \boldsymbol{A})~da
</math>
Using the assumption that the curl of <math>\boldsymbol{A}</math> is zero, we get
:<math>
  \oint_{\partial\Omega} \boldsymbol{A}~ds = 0 \quad \implies \quad
  \int_{AB} \boldsymbol{A}\cdot d\mathbf{X} + \int_{BA} \boldsymbol{A}\cdot d\mathbf{X} = 0
</math>
Hence the integral is path independent and the compatibility condition is sufficient to ensure a unique <math>\mathbf{v}</math> field, provided that the body is simply connected.
 
== Compatibility of the deformation gradient ==
The compatibility condition for the deformation gradient is obtained directly from the above proof by observing that
:<math>
  \boldsymbol{F} = \cfrac{\partial \mathbf{x}}{\partial \mathbf{X}} = \boldsymbol{\nabla}\mathbf{x}
</math>
Then the necessary and sufficient conditions for the existence of a compatible <math>\boldsymbol{F}</math> field over a simply connected body are
:<math>
  \boldsymbol{\nabla}\times\boldsymbol{F} = \boldsymbol{0}
</math>
 
==Compatibility of infinitesimal strains==
The compatibility problem for small strains can be stated as follows.
 
Given a symmetric second order tensor field <math>\boldsymbol{\epsilon}</math> when is it possible to construct a vector field <math>\mathbf{u}</math> such that
:<math>
  \boldsymbol{\epsilon} = \frac{1}{2} [\boldsymbol{\nabla}\mathbf{u} + (\boldsymbol{\nabla}\mathbf{u})^T]
</math>
 
===Necessary conditions===
Suppose that there exists <math>\mathbf{u}</math> such that the expression for <math>\boldsymbol{\epsilon}</math> holds. Now
:<math>
  \boldsymbol{\nabla}\mathbf{u} = \boldsymbol{\epsilon} + \boldsymbol{\omega}
</math>
where
:<math>
  \boldsymbol{\omega} := \frac{1}{2} [\boldsymbol{\nabla}\mathbf{u} - (\boldsymbol{\nabla}\mathbf{u})^T]
</math>
Therefore, in index notation,
:<math>
  \boldsymbol{\nabla} \boldsymbol{\omega} \equiv \omega_{ij,k} = \frac{1}{2} (u_{i,jk} - u_{j,ik}) = \frac{1}{2} (u_{i,jk} + u_{k,ji} - u_{j,ik} - u_{k,ji}) = \varepsilon_{ik,j} - \varepsilon_{jk,i}
</math>
If <math>\boldsymbol{\omega}</math> is continuously differentiable we have <math>\omega_{ij,kl} = \omega_{ij,lk}</math>. Hence,
:<math>
  \varepsilon_{ik,jl} - \varepsilon_{jk,il} - \varepsilon_{il,jk} + \varepsilon_{jl,ik} = 0
</math>
In direct tensor notation
:<math>
  \boldsymbol{\nabla} \times (\boldsymbol{\nabla} \times\boldsymbol{\epsilon}) = \boldsymbol{0}
</math>
The above are necessary conditions.  If <math>\mathbf{w}</math> is the [[infinitesimal strain theory|infinitesimal rotation vector]] then <math>\boldsymbol{\nabla} \times \boldsymbol{\epsilon} = \boldsymbol{\nabla} \mathbf{w}</math>. Hence the necessary condition may also be written as <math>\boldsymbol{\nabla} \times (\boldsymbol{\nabla \mathbf{w})} = \boldsymbol{0}</math>.
 
===Sufficient conditions===
Let us now assume that the condition <math>\boldsymbol{\nabla} \times (\boldsymbol{\nabla} \times\boldsymbol{\epsilon}) = \boldsymbol{0}</math> is satisfied in a portion of a body. Is this condition sufficient to guarantee the existence of a continuous, single-valued displacement field <math>\mathbf{u}</math>?
 
The first step in the process is to show that this condition implies that the [[infinitesimal strain theory|infinitesimal rotation tensor]] <math>\boldsymbol{\omega}</math> is uniquely defined.  To do that we integrate <math>\boldsymbol{\nabla} \mathbf{w}</math> along the path <math>\mathbf{X}_A</math> to <math>\mathbf{X}_B</math>, i.e.,
:<math>
  \mathbf{w}(\mathbf{X}_B) - \mathbf{w}(\mathbf{X}_A) = \int_{\mathbf{X}_A}^{\mathbf{X}_B} \boldsymbol{\nabla} \mathbf{w}\cdot d\mathbf{X}
    = \int_{\mathbf{X}_A}^{\mathbf{X}_B} (\boldsymbol{\nabla} \times \boldsymbol{\epsilon})\cdot d\mathbf{X}
</math>
Note that we need to know  a reference <math>\mathbf{w}(\mathbf{X}_A)</math> to fix the rigid body rotation. The field <math>\mathbf{w}(\mathbf{X})</math> is uniquely determined only if the contour integral along a closed contour between <math>\mathbf{X}_A</math> and <math>\mathbf{X}_b</math> is zero, i.e.,
:<math>
  \oint_{\mathbf{X}_A}^{\mathbf{X}_B} (\boldsymbol{\nabla} \times \boldsymbol{\epsilon})\cdot d\mathbf{X}  = \boldsymbol{0}
</math>
But from Stokes' theorem for a simply-connected body and the necessary condition for compatibility
:<math>
  \oint_{\mathbf{X}_A}^{\mathbf{X}_B} (\boldsymbol{\nabla} \times \boldsymbol{\epsilon})\cdot d\mathbf{X}  = \int_{\Omega_{AB}} \mathbf{n}\cdot(\boldsymbol{\nabla} \times \boldsymbol{\nabla}\times\boldsymbol{\epsilon})~da
    = \boldsymbol{0}
</math>
Therefore the field <math>\mathbf{w}</math> is uniquely defined which implies that the infinitesimal rotation tensor <math>\boldsymbol{\omega}</math> is also uniquely defined, provided the body is simply connected.
 
In the next step of the process we will consider the uniqueness of the displacement field <math>\mathbf{u}</math>.  As before we integrate the displacement gradient
:<math>
  \mathbf{u}(\mathbf{X}_B) - \mathbf{u}(\mathbf{X}_A) = \int_{\mathbf{X}_A}^{\mathbf{X}_B} \boldsymbol{\nabla} \mathbf{u}\cdot d\mathbf{X}
    = \int_{\mathbf{X}_A}^{\mathbf{X}_B} (\boldsymbol{\epsilon} + \boldsymbol{\omega})\cdot d\mathbf{X}
</math>
From Stokes' theorem and using the relations <math>\boldsymbol{\nabla} \times \boldsymbol{\epsilon} = \boldsymbol{\nabla} \mathbf{w} = -\boldsymbol{\nabla} \times \Omega</math> we have
:<math>
  \oint_{\mathbf{X}_A}^{\mathbf{X}_B} (\boldsymbol{\epsilon} + \boldsymbol{\omega})\cdot d\mathbf{X} = \int_{\Omega_{AB}} \mathbf{n}\cdot(\boldsymbol{\nabla} \times \boldsymbol{\epsilon}+\boldsymbol{\nabla} \times \boldsymbol{\omega})~da = \boldsymbol{0}
</math>
Hence the displacement field <math>\mathbf{u}</math> is also determined uniquely.  Hence the compatibility conditions are sufficient to guarantee the existence of a unique displacement field <math>\mathbf{u}</math> in a simply-connected body.
 
==Compatibility for Right Cauchy-Green Deformation field==
The compatibility problem for the Right Cauchy-Green deformation field can be posed as follows.
 
''' Problem:''' Let <math>\boldsymbol{C}(\mathbf{X})</math> be a positive definite symmetric tensor field defined on the reference configuration.  Under what conditions on <math>\boldsymbol{C}</math> does there exist a deformed configuration marked by the position field <math>\mathbf{x}(\mathbf{X})</math> such that
:<math>
  (1)\quad\left(\frac{\partial \mathbf{x}}{\partial \mathbf{X}}\right)^T\cdot\left(\frac{\partial \mathbf{x}}{\partial \mathbf{X}}\right) = \boldsymbol{C}
</math>
 
===Necessary conditions===
Suppose that a field <math>\mathbf{x}(\mathbf{X})</math> exists that satisfies condition (1).  In terms of components with respect to a rectangular Cartesian basis
:<math>
  \frac{\partial x^i}{\partial X^\alpha}\frac{\partial x^i}{\partial X^\beta} = C_{\alpha\beta}
</math>
From [[finite strain theory]] we know that <math>C_{\alpha\beta} = g_{\alpha\beta}</math>.  Hence we can write
:<math>
  \delta_{ij}~\frac{\partial x^i}{\partial X^\alpha}~\frac{\partial x^j}{\partial X^\beta} = g_{\alpha\beta}
</math>
For two symmetric second-order tensor field that are mapped one-to-one we also have the [[Finite_strain_theory#Some_relations_between_deformation_measures_and_Christoffel_symbols|relation]]
:<math>
  G_{ij} = \frac{\partial X^\alpha}{\partial x^i}~\frac{\partial X^\beta}{\partial x^j}~g_{\alpha\beta}
</math>
From the relation between of <math>G_{ij}</math> and <math>g_{\alpha\beta}</math> that <math>\delta_{ij} = G_{ij}</math>, we have
:<math>
  _{(x)}\Gamma_{ij}^k = 0
</math>
Then, from the relation
:<math>
\frac{\partial^2 x^m}{\partial X^\alpha \partial X^\beta}  = \frac{\partial x^m}{\partial X^\mu}\,_{(X)}\Gamma^\mu_{\alpha\beta} - \frac{\partial x^i}{\partial X^\alpha}~\frac{\partial x^j}{\partial X^\beta} \,_{(x)}\Gamma^m_{ij}
</math>
we have
:<math>
\frac{\partial F^m_{~\alpha}}{\partial X^\beta}  = F^m_{~\mu}\,_{(X)}\Gamma^\mu_{\alpha\beta} \qquad; ~~
  F^i_{~\alpha} := \frac{\partial x^i}{\partial X^\alpha}
</math>
From [[Finite_strain_theory#Some_relations_between_deformation_measures_and_Christoffel_symbols|finite strain theory]] we also have
:<math>
  _{(X)}\Gamma_{\alpha\beta\gamma} = \frac{1}{2}\left(\frac{\partial g_{\alpha\gamma}}{\partial X^\beta} + \frac{\partial g_{\beta\gamma}}{\partial X^\alpha} - \frac{\partial g_{\alpha\beta}}{\partial X^\gamma}\right) ~;~~
  _{(X)}\Gamma^\nu_{\alpha\beta} = g^{\nu\gamma} \,_{(X)}\Gamma_{\alpha\beta\gamma} ~;~~
  g_{\alpha\beta} = C_{\alpha\beta} ~;~~ g^{\alpha\beta} = C^{\alpha\beta}
</math>
Therefore
:<math>
\,_{(X)}\Gamma^\mu_{\alpha\beta} = \cfrac{C^{\mu\gamma}}{2}\left(\frac{\partial C_{\alpha\gamma}}{\partial X^\beta} + \frac{\partial C_{\beta\gamma}}{\partial X^\alpha} - \frac{\partial C_{\alpha\beta}}{\partial X^\gamma}\right)
</math>
and we have
:<math>
\frac{\partial F^m_{~\alpha}}{\partial X^\beta}  = F^m_{~\mu}~\cfrac{C^{\mu\gamma}}{2}\left(\frac{\partial C_{\alpha\gamma}}{\partial X^\beta} + \frac{\partial C_{\beta\gamma}}{\partial X^\alpha} - \frac{\partial C_{\alpha\beta}}{\partial X^\gamma}\right)
</math>
Again, using the commutative nature of the order of differentiation, we have
:<math>
\frac{\partial^2 F^m_{~\alpha}}{\partial X^\beta \partial X^\rho} = \frac{\partial^2 F^m_{~\alpha}}{\partial X^\rho \partial X^\beta} 
\implies
  \frac{\partial F^m_{~\mu}}{\partial X^\rho}\,_{(X)}\Gamma^\mu_{\alpha\beta} +
  F^m_{~\mu}~\frac{\partial }{\partial X^\rho}[\,_{(X)}\Gamma^\mu_{\alpha\beta}] =
  \frac{\partial F^m_{~\mu}}{\partial X^\beta}\,_{(X)}\Gamma^\mu_{\alpha\rho} +
  F^m_{~\mu}~\frac{\partial }{\partial X^\beta}[\,_{(X)}\Gamma^\mu_{\alpha\rho}]
</math>
or
:<math>
  F^m_{~\gamma}\,_{(X)}\Gamma^\gamma_{\mu\rho}\,_{(X)}\Gamma^\mu_{\alpha\beta} +
  F^m_{~\mu}~\frac{\partial }{\partial X^\rho}[\,_{(X)}\Gamma^\mu_{\alpha\beta}] =
  F^m_{~\gamma}\,_{(X)}\Gamma^\gamma_{\mu\beta}\,_{(X)}\Gamma^\mu_{\alpha\rho} +
  F^m_{~\mu}~\frac{\partial }{\partial X^\beta}[\,_{(X)}\Gamma^\mu_{\alpha\rho}]
</math>
After collecting terms we get
:<math>
  F^m_{~\gamma}\left(\,_{(X)}\Gamma^\gamma_{\mu\rho}\,_{(X)}\Gamma^\mu_{\alpha\beta} +
  \frac{\partial }{\partial X^\rho}[\,_{(X)}\Gamma^\gamma_{\alpha\beta}] -
  \,_{(X)}\Gamma^\gamma_{\mu\beta}\,_{(X)}\Gamma^\mu_{\alpha\rho} -
  \frac{\partial }{\partial X^\beta}[\,_{(X)}\Gamma^\gamma_{\alpha\rho}]\right) = 0
</math>
From the definition of <math>F^m_{\gamma}</math> we observe that it is invertible and hence cannot be zero. Therefore,
:<math>
  R^\gamma_{\alpha\beta\rho} :=
  \frac{\partial }{\partial X^\rho}[\,_{(X)}\Gamma^\gamma_{\alpha\beta}] -
  \frac{\partial }{\partial X^\beta}[\,_{(X)}\Gamma^\gamma_{\alpha\rho}] +
  \,_{(X)}\Gamma^\gamma_{\mu\rho}\,_{(X)}\Gamma^\mu_{\alpha\beta} -
  \,_{(X)}\Gamma^\gamma_{\mu\beta}\,_{(X)}\Gamma^\mu_{\alpha\rho} = 0
</math>
We can show these are the mixed components of the [[Riemann-Christoffel curvature tensor]]. Therefore the necessary conditions for <math>\boldsymbol{C}</math>-compatibility are that the Riemann-Christoffel curvature of the deformation is zero.
 
===Sufficient conditions===
The proof of sufficiency is a bit more involved.<ref name="acharya"/><ref name=Blume>Blume, J. A., 1989, "Compatibility conditions for a left Cauchy-Green strain field", J. Elasticity, v. 21, p. 271-308.</ref>  We start with the assumption that
:<math>
  R^\gamma_{\alpha\beta\rho} = 0 ~;~~ g_{\alpha\beta} = C_{\alpha\beta}
</math>
We have to show that there exist <math>\mathbf{x}</math> and <math>\mathbf{X}</math> such that
:<math>
  \frac{\partial x^i}{\partial X^\alpha}\frac{\partial x^i}{\partial X^\beta} = C_{\alpha\beta}
</math>
From a theorem by T.Y.Thomas <ref name=Thomas>Thomas, T. Y., 1934, "Systems of total differential equations defined over simply connected domains", Annals of Mathematics, 35(4), p. 930-734</ref> we know that the system of equations
:<math>
  \frac{\partial F^i_{~\alpha}}{\partial X^\beta} = F^i_{~\gamma}~\,_{(X)}\Gamma^\gamma_{\alpha\beta}
</math>
has unique solutions <math>F^i_{~\alpha}</math> over simply connected domains if
:<math>
  _{(X)}\Gamma^\gamma_{\alpha\beta} = _{(X)}\Gamma^\gamma_{\beta\alpha} ~;~~
  R^\gamma_{\alpha\beta\rho} = 0
</math>
The first of these is true from the defining of <math>\Gamma^i_{jk}</math> and the second is assumed. Hence the assumed condition gives us a unique <math>F^i_{~\alpha}</math> that is <math>C^2</math> continuous. 
 
Next consider the system of equations
:<math>
  \frac{\partial x^i}{\partial X^\alpha} = F^i_{~\alpha}
</math>
Since <math>F^i_{~\alpha}</math> is <math>C^2</math> and the body is simply connected there exists some solution <math>x^i(X^\alpha)</math> to the above equations. We can show that the <math>x^i</math> also satisfy the property that
:<math>
  \det\left|\frac{\partial x^i}{\partial X^\alpha}\right| \ne 0
</math>
We can also show that the relation
:<math>
  \frac{\partial x^i}{\partial X^\alpha}~g^{\alpha\beta}~\frac{\partial x^j}{\partial X^\beta} = \delta^{ij}
</math>
implies that
:<math>
  g_{\alpha\beta} = C_{\alpha\beta} = \frac{\partial x^k}{\partial X^\alpha}~\frac{\partial x^k}{\partial X^\beta}
</math>
If we associate these quantities with tensor fields we can show that <math>\frac{\partial \mathbf{x}}{\partial \mathbf{X}}</math> is invertible and the constructed tensor field satisfies the expression for <math>\boldsymbol{C}</math>.
 
== See also ==
* [[Saint-Venant's compatibility condition]]
* [[Linear elasticity]]
* [[Deformation (mechanics)]]
* [[Infinitesimal strain theory]]
* [[Finite strain theory]]
* [[Tensor derivative (continuum mechanics)]]
* [[Curvilinear coordinates]]
 
== References ==
<references />
 
== External links ==
*[http://www.imechanica.org/node/3786 Prof. Amit Acharya's notes on compatibility on iMechanica]
*[http://www.ce.berkeley.edu/~coby/plas/pdf/book.pdf Plasticity by J. Lubliner, sec. 1.2.4 p. 35]
 
[[Category:Continuum mechanics]]
[[Category:Elasticity (physics)]]

Latest revision as of 20:18, 26 November 2014


There are many types of bike brakes for different applications. Various means of transportation come second - from yachts or boats and cars to motorcycles. The contemporary bikes for mountains are provided with stronger and lighter frame types in addition to pioneering form and design. Mountain bike frames spend a lot of time on roads these days, too. If you've ridden for years and years and know what you need, buying online does make sense.

Once you are sure that this is what you want to do, it is now time for you to decide on the bike. Once you're comfortable coasting, dropping, standing, pedaling, spinning, and switching gears, you'll be ready to hit the trails, and tackle any challenge along the way. So how exactly would you go about making this dream a reality. If biking is your passion or your favorite sport, then you know the worth of good accessories while riding. If you are riding through a rock garden, for instance, you don't want a fork that is bouncing you all over the place.

Some also use the liner as a size aid, for example BMX helmets can come as one-size, with optional liner kits to change the helmets size. An Indian travois style trailer is very simple to make and very effective for heavy loads. The whole set-up, including your bag, should weigh no more than 20kgs (44lbs). Another way to classify brakes is by mounting style. The most common sized crankset is the 170 millimeter.

There are still other types of bikes that require different types of tires. This article discusses how to go about buying a mountain bike. Just remember your heart is a muscle and can be developed the more you practice. With a Shimano 105 groupset and FSA Gossamer chain set, it's a great choice for under. I've never been into cycling and wanted to give it a try for exercise.

If you have any kind of concerns relating to where and the best ways to utilize Choosing the right ride for you mountain bike sizing., you could call us at our own site. I ran with Hale and we also tried this method, it was effective. Instability: You are much more likely to lose your balance while on a big bike. There are good bikes available at reasonable prices, and there are some overpriced stinkers. Suspension: Road bikes are built with a sole purpose of providing greater speed; they do not possess this feature, although they have certain materials which absorb the shocks of the uneven roads. They will provide you with a honest price, yet be ready to pay in between $650 and $4,000 on dual suspensions and in between $470 and $670 for hardtails.