fbpx
Wikipedia

Gradient theorem

The gradient theorem, also known as the fundamental theorem of calculus for line integrals, says that a line integral through a gradient field can be evaluated by evaluating the original scalar field at the endpoints of the curve. The theorem is a generalization of the second fundamental theorem of calculus to any curve in a plane or space (generally n-dimensional) rather than just the real line.

For φ : URnR as a differentiable function and γ as any continuous curve in U which starts at a point p and ends at a point q, then

where φ denotes the gradient vector field of φ.

The gradient theorem implies that line integrals through gradient fields are path-independent. In physics this theorem is one of the ways of defining a conservative force. By placing φ as potential, φ is a conservative field. Work done by conservative forces does not depend on the path followed by the object, but only the end points, as the above equation shows.

The gradient theorem also has an interesting converse: any path-independent vector field can be expressed as the gradient of a scalar field. Just like the gradient theorem itself, this converse has many striking consequences and applications in both pure and applied mathematics.

Proof edit

If φ is a differentiable function from some open subset URn to R and r is a differentiable function from some closed interval [a, b] to U (Note that r is differentiable at the interval endpoints a and b. To do this, r is defined on an interval that is larger than and includes [a, b].), then by the multivariate chain rule, the composite function φr is differentiable on [a, b]:

 

for all t in [a, b]. Here the denotes the usual inner product.

Now suppose the domain U of φ contains the differentiable curve γ with endpoints p and q. (This is oriented in the direction from p to q). If r parametrizes γ for t in [a, b] (i.e., r represents γ as a function of t), then

 

where the definition of a line integral is used in the first equality, the above equation is used in the second equality, and the second fundamental theorem of calculus is used in the third equality.[1]

Even if the gradient theorem (also called fundamental theorem of calculus for line integrals) has been proved for a differentiable (so looked as smooth) curve so far, the theorem is also proved for a piecewise-smooth curve since this curve is made by joining multiple differentiable curves so the proof for this curve is made by the proof per differentiable curve component.[2]

Examples edit

Example 1 edit

Suppose γR2 is the circular arc oriented counterclockwise from (5, 0) to (−4, 3). Using the definition of a line integral,

 

This result can be obtained much more simply by noticing that the function   has gradient  , so by the Gradient Theorem:

 

Example 2 edit

For a more abstract example, suppose γRn has endpoints p, q, with orientation from p to q. For u in Rn, let |u| denote the Euclidean norm of u. If α ≥ 1 is a real number, then

 

Here the final equality follows by the gradient theorem, since the function f(x) = |x|α+1 is differentiable on Rn if α ≥ 1.

If α < 1 then this equality will still hold in most cases, but caution must be taken if γ passes through or encloses the origin, because the integrand vector field |x|α − 1x will fail to be defined there. However, the case α = −1 is somewhat different; in this case, the integrand becomes |x|−2x = ∇(log |x|), so that the final equality becomes log |q| − log |p|.

Note that if n = 1, then this example is simply a slight variant of the familiar power rule from single-variable calculus.

Example 3 edit

Suppose there are n point charges arranged in three-dimensional space, and the i-th point charge has charge Qi and is located at position pi in R3. We would like to calculate the work done on a particle of charge q as it travels from a point a to a point b in R3. Using Coulomb's law, we can easily determine that the force on the particle at position r will be

 

Here |u| denotes the Euclidean norm of the vector u in R3, and k = 1/(4πε0), where ε0 is the vacuum permittivity.

Let γR3 − {p1, ..., pn} be an arbitrary differentiable curve from a to b. Then the work done on the particle is

 

Now for each i, direct computation shows that

 

Thus, continuing from above and using the gradient theorem,

 

We are finished. Of course, we could have easily completed this calculation using the powerful language of electrostatic potential or electrostatic potential energy (with the familiar formulas W = −ΔU = −qΔV). However, we have not yet defined potential or potential energy, because the converse of the gradient theorem is required to prove that these are well-defined, differentiable functions and that these formulas hold (see below). Thus, we have solved this problem using only Coulomb's Law, the definition of work, and the gradient theorem.

Converse of the gradient theorem edit

The gradient theorem states that if the vector field F is the gradient of some scalar-valued function (i.e., if F is conservative), then F is a path-independent vector field (i.e., the integral of F over some piecewise-differentiable curve is dependent only on end points). This theorem has a powerful converse:

Theorem —  If F is a path-independent vector field, then F is the gradient of some scalar-valued function.[3]

It is straightforward to show that a vector field is path-independent if and only if the integral of the vector field over every closed loop in its domain is zero. Thus the converse can alternatively be stated as follows: If the integral of F over every closed loop in the domain of F is zero, then F is the gradient of some scalar-valued function.

Proof of the converse edit

Suppose U is an open, path-connected subset of Rn, and F : URn is a continuous and path-independent vector field. Fix some element a of U, and define f : UR by

 
Here γ[a, x] is any (differentiable) curve in U originating at a and terminating at x. We know that f is well-defined because F is path-independent.

Let v be any nonzero vector in Rn. By the definition of the directional derivative,

 
To calculate the integral within the final limit, we must parametrize γ[x, x + tv]. Since F is path-independent, U is open, and t is approaching zero, we may assume that this path is a straight line, and parametrize it as u(s) = x + sv for 0 < s < t. Now, since u'(s) = v, the limit becomes
 
where the first equality is from the definition of the derivative with a fact that the integral is equal to 0 at t = 0, and the second equality is from the first fundamental theorem of calculus. Thus we have a formula for vf, (one of ways to represent the directional derivative) where v is arbitrary; for   (see its full definition above), its directional derivative with respect to v is
 
where the first two equalities just show different representations of the directional derivative. According to the definition of the gradient of a scalar function f,  , thus we have found a scalar-valued function f whose gradient is the path-independent vector field F (i.e., F is a conservative vector field.), as desired.[3]

Example of the converse principle edit

To illustrate the power of this converse principle, we cite an example that has significant physical consequences. In classical electromagnetism, the electric force is a path-independent force; i.e. the work done on a particle that has returned to its original position within an electric field is zero (assuming that no changing magnetic fields are present).

Therefore, the above theorem implies that the electric force field Fe : SR3 is conservative (here S is some open, path-connected subset of R3 that contains a charge distribution). Following the ideas of the above proof, we can set some reference point a in S, and define a function Ue: SR by

 

Using the above proof, we know Ue is well-defined and differentiable, and Fe = −∇Ue (from this formula we can use the gradient theorem to easily derive the well-known formula for calculating work done by conservative forces: W = −ΔU). This function Ue is often referred to as the electrostatic potential energy of the system of charges in S (with reference to the zero-of-potential a). In many cases, the domain S is assumed to be unbounded and the reference point a is taken to be "infinity", which can be made rigorous using limiting techniques. This function Ue is an indispensable tool used in the analysis of many physical systems.

Generalizations edit

Many of the critical theorems of vector calculus generalize elegantly to statements about the integration of differential forms on manifolds. In the language of differential forms and exterior derivatives, the gradient theorem states that

 

for any 0-form, ϕ, defined on some differentiable curve γRn (here the integral of ϕ over the boundary of the γ is understood to be the evaluation of ϕ at the endpoints of γ).

Notice the striking similarity between this statement and the generalized Stokes’ theorem, which says that the integral of any compactly supported differential form ω over the boundary of some orientable manifold Ω is equal to the integral of its exterior derivative dω over the whole of Ω, i.e.,

 

This powerful statement is a generalization of the gradient theorem from 1-forms defined on one-dimensional manifolds to differential forms defined on manifolds of arbitrary dimension.

The converse statement of the gradient theorem also has a powerful generalization in terms of differential forms on manifolds. In particular, suppose ω is a form defined on a contractible domain, and the integral of ω over any closed manifold is zero. Then there exists a form ψ such that ω = dψ. Thus, on a contractible domain, every closed form is exact. This result is summarized by the Poincaré lemma.

See also edit

References edit

  1. ^ Williamson, Richard and Trotter, Hale. (2004). Multivariable Mathematics, Fourth Edition, p. 374. Pearson Education, Inc.
  2. ^ Stewart, James (2015). "16.3 The Fundamental Theorem for Line Integrals". Calculus (8th ed.). Cengage Learning. pp. 1127–1128. ISBN 978-1-285-74062-1.
  3. ^ a b "Williamson, Richard and Trotter, Hale. (2004). Multivariable Mathematics, Fourth Edition, p. 410. Pearson Education, Inc."

gradient, theorem, gradient, theorem, also, known, fundamental, theorem, calculus, line, integrals, says, that, line, integral, through, gradient, field, evaluated, evaluating, original, scalar, field, endpoints, curve, theorem, generalization, second, fundame. The gradient theorem also known as the fundamental theorem of calculus for line integrals says that a line integral through a gradient field can be evaluated by evaluating the original scalar field at the endpoints of the curve The theorem is a generalization of the second fundamental theorem of calculus to any curve in a plane or space generally n dimensional rather than just the real line For f U Rn R as a differentiable function and g as any continuous curve in U which starts at a point p and ends at a point q then g f r d r f q f p displaystyle int gamma nabla varphi mathbf r cdot mathrm d mathbf r varphi left mathbf q right varphi left mathbf p right where f denotes the gradient vector field of f The gradient theorem implies that line integrals through gradient fields are path independent In physics this theorem is one of the ways of defining a conservative force By placing f as potential f is a conservative field Work done by conservative forces does not depend on the path followed by the object but only the end points as the above equation shows The gradient theorem also has an interesting converse any path independent vector field can be expressed as the gradient of a scalar field Just like the gradient theorem itself this converse has many striking consequences and applications in both pure and applied mathematics Contents 1 Proof 2 Examples 2 1 Example 1 2 2 Example 2 2 3 Example 3 3 Converse of the gradient theorem 3 1 Proof of the converse 3 2 Example of the converse principle 4 Generalizations 5 See also 6 ReferencesProof editIf f is a differentiable function from some open subset U Rn to R and r is a differentiable function from some closed interval a b to U Note that r is differentiable at the interval endpoints a and b To do this r is defined on an interval that is larger than and includes a b then by the multivariate chain rule the composite function f r is differentiable on a b d d t f r t f r t r t displaystyle frac mathrm d mathrm d t varphi circ mathbf r t nabla varphi mathbf r t cdot mathbf r t nbsp for all t in a b Here the denotes the usual inner product Now suppose the domain U of f contains the differentiable curve g with endpoints p and q This is oriented in the direction from p to q If r parametrizes g for t in a b i e r represents g as a function of t then g f r d r a b f r t r t d t a b d d t f r t d t f r b f r a f q f p displaystyle begin aligned int gamma nabla varphi mathbf r cdot mathrm d mathbf r amp int a b nabla varphi mathbf r t cdot mathbf r t mathrm d t amp int a b frac d dt varphi mathbf r t mathrm d t varphi mathbf r b varphi mathbf r a varphi left mathbf q right varphi left mathbf p right end aligned nbsp where the definition of a line integral is used in the first equality the above equation is used in the second equality and the second fundamental theorem of calculus is used in the third equality 1 Even if the gradient theorem also called fundamental theorem of calculus for line integrals has been proved for a differentiable so looked as smooth curve so far the theorem is also proved for a piecewise smooth curve since this curve is made by joining multiple differentiable curves so the proof for this curve is made by the proof per differentiable curve component 2 Examples editExample 1 edit Suppose g R2 is the circular arc oriented counterclockwise from 5 0 to 4 3 Using the definition of a line integral g y d x x d y 0 p tan 1 3 4 5 sin t 5 sin t 5 cos t 5 cos t d t 0 p tan 1 3 4 25 sin 2 t cos 2 t d t 0 p tan 1 3 4 25 cos 2 t d t 25 2 sin 2 t 0 p tan 1 3 4 25 2 sin 2 p 2 tan 1 3 4 25 2 sin 2 tan 1 3 4 25 3 4 3 4 2 1 12 displaystyle begin aligned int gamma y mathrm d x x mathrm d y amp int 0 pi tan 1 left frac 3 4 right 5 sin t 5 sin t 5 cos t 5 cos t mathrm d t amp int 0 pi tan 1 left frac 3 4 right 25 left sin 2 t cos 2 t right mathrm d t amp int 0 pi tan 1 left frac 3 4 right 25 cos 2t mathrm d t left tfrac 25 2 sin 2t right 0 pi tan 1 left tfrac 3 4 right 5em amp tfrac 25 2 sin left 2 pi 2 tan 1 left tfrac 3 4 right right 5em amp tfrac 25 2 sin left 2 tan 1 left tfrac 3 4 right right frac 25 3 4 3 4 2 1 12 end aligned nbsp This result can be obtained much more simply by noticing that the function f x y x y displaystyle f x y xy nbsp has gradient f x y y x displaystyle nabla f x y y x nbsp so by the Gradient Theorem g y d x x d y g x y d x d y x y 5 0 4 3 4 3 5 0 12 displaystyle int gamma y mathrm d x x mathrm d y int gamma nabla xy cdot mathrm d x mathrm d y xy 5 0 4 3 4 cdot 3 5 cdot 0 12 nbsp Example 2 edit For a more abstract example suppose g Rn has endpoints p q with orientation from p to q For u in Rn let u denote the Euclidean norm of u If a 1 is a real number then g x a 1 x d x 1 a 1 g a 1 x a 1 2 x d x 1 a 1 g x a 1 d x q a 1 p a 1 a 1 displaystyle begin aligned int gamma mathbf x alpha 1 mathbf x cdot mathrm d mathbf x amp frac 1 alpha 1 int gamma alpha 1 mathbf x alpha 1 2 mathbf x cdot mathrm d mathbf x amp frac 1 alpha 1 int gamma nabla mathbf x alpha 1 cdot mathrm d mathbf x frac mathbf q alpha 1 mathbf p alpha 1 alpha 1 end aligned nbsp Here the final equality follows by the gradient theorem since the function f x x a 1 is differentiable on Rn if a 1 If a lt 1 then this equality will still hold in most cases but caution must be taken if g passes through or encloses the origin because the integrand vector field x a 1x will fail to be defined there However the case a 1 is somewhat different in this case the integrand becomes x 2x log x so that the final equality becomes log q log p Note that if n 1 then this example is simply a slight variant of the familiar power rule from single variable calculus Example 3 edit Suppose there are n point charges arranged in three dimensional space and the i th point charge has charge Qi and is located at position pi in R3 We would like to calculate the work done on a particle of charge q as it travels from a point a to a point b in R3 Using Coulomb s law we can easily determine that the force on the particle at position r will beF r k q i 1 n Q i r p i r p i 3 displaystyle mathbf F mathbf r kq sum i 1 n frac Q i mathbf r mathbf p i left mathbf r mathbf p i right 3 nbsp Here u denotes the Euclidean norm of the vector u in R3 and k 1 4pe0 where e0 is the vacuum permittivity Let g R3 p1 pn be an arbitrary differentiable curve from a to b Then the work done on the particle isW g F r d r g k q i 1 n Q i r p i r p i 3 d r k q i 1 n Q i g r p i r p i 3 d r displaystyle W int gamma mathbf F mathbf r cdot mathrm d mathbf r int gamma left kq sum i 1 n frac Q i mathbf r mathbf p i left mathbf r mathbf p i right 3 right cdot mathrm d mathbf r kq sum i 1 n left Q i int gamma frac mathbf r mathbf p i left mathbf r mathbf p i right 3 cdot mathrm d mathbf r right nbsp Now for each i direct computation shows thatr p i r p i 3 1 r p i displaystyle frac mathbf r mathbf p i left mathbf r mathbf p i right 3 nabla frac 1 left mathbf r mathbf p i right nbsp Thus continuing from above and using the gradient theorem W k q i 1 n Q i g 1 r p i d r k q i 1 n Q i 1 a p i 1 b p i displaystyle W kq sum i 1 n left Q i int gamma nabla frac 1 left mathbf r mathbf p i right cdot mathrm d mathbf r right kq sum i 1 n Q i left frac 1 left mathbf a mathbf p i right frac 1 left mathbf b mathbf p i right right nbsp We are finished Of course we could have easily completed this calculation using the powerful language of electrostatic potential or electrostatic potential energy with the familiar formulas W DU qDV However we have not yet defined potential or potential energy because the converse of the gradient theorem is required to prove that these are well defined differentiable functions and that these formulas hold see below Thus we have solved this problem using only Coulomb s Law the definition of work and the gradient theorem Converse of the gradient theorem editThe gradient theorem states that if the vector field F is the gradient of some scalar valued function i e if F is conservative then F is a path independent vector field i e the integral of F over some piecewise differentiable curve is dependent only on end points This theorem has a powerful converse Theorem If F is a path independent vector field then F is the gradient of some scalar valued function 3 It is straightforward to show that a vector field is path independent if and only if the integral of the vector field over every closed loop in its domain is zero Thus the converse can alternatively be stated as follows If the integral of F over every closed loop in the domain of F is zero then F is the gradient of some scalar valued function Proof of the converse edit Suppose U is an open path connected subset of Rn and F U Rn is a continuous and path independent vector field Fix some element a of U and define f U R byf x g a x F u d u displaystyle f mathbf x int gamma mathbf a mathbf x mathbf F mathbf u cdot mathrm d mathbf u nbsp Here g a x is any differentiable curve in U originating at a and terminating at x We know that f is well defined because F is path independent Let v be any nonzero vector in Rn By the definition of the directional derivative f x v lim t 0 f x t v f x t lim t 0 g a x t v F u d u g a x F u d u t lim t 0 1 t g x x t v F u d u displaystyle begin aligned frac partial f mathbf x partial mathbf v amp lim t to 0 frac f mathbf x t mathbf v f mathbf x t amp lim t to 0 frac int gamma mathbf a mathbf x t mathbf v mathbf F mathbf u cdot mathrm d mathbf u int gamma mathbf a mathbf x mathbf F mathbf u cdot d mathbf u t amp lim t to 0 frac 1 t int gamma mathbf x mathbf x t mathbf v mathbf F mathbf u cdot mathrm d mathbf u end aligned nbsp To calculate the integral within the final limit we must parametrize g x x tv Since F is path independent U is open and t is approaching zero we may assume that this path is a straight line and parametrize it as u s x sv for 0 lt s lt t Now since u s v the limit becomeslim t 0 1 t 0 t F u s u s d s d d t 0 t F x s v v d s t 0 F x v displaystyle lim t to 0 frac 1 t int 0 t mathbf F mathbf u s cdot mathbf u s mathrm d s frac mathrm d mathrm d t int 0 t mathbf F mathbf x s mathbf v cdot mathbf v mathrm d s bigg t 0 mathbf F mathbf x cdot mathbf v nbsp where the first equality is from the definition of the derivative with a fact that the integral is equal to 0 at t 0 and the second equality is from the first fundamental theorem of calculus Thus we have a formula for vf one of ways to represent the directional derivative where v is arbitrary for f x g a x F u d u displaystyle f mathbf x int gamma mathbf a mathbf x mathbf F mathbf u cdot mathrm d mathbf u nbsp see its full definition above its directional derivative with respect to v is f x v v f x D v f x F x v displaystyle frac partial f mathbf x partial mathbf v partial mathbf v f mathbf x D mathbf v f mathbf x mathbf F mathbf x cdot mathbf v nbsp where the first two equalities just show different representations of the directional derivative According to the definition of the gradient of a scalar function f f x F x displaystyle nabla f mathbf x mathbf F mathbf x nbsp thus we have found a scalar valued function f whose gradient is the path independent vector field F i e F is a conservative vector field as desired 3 Example of the converse principle edit Main article Electric potential energy To illustrate the power of this converse principle we cite an example that has significant physical consequences In classical electromagnetism the electric force is a path independent force i e the work done on a particle that has returned to its original position within an electric field is zero assuming that no changing magnetic fields are present Therefore the above theorem implies that the electric force field Fe S R3 is conservative here S is some open path connected subset of R3 that contains a charge distribution Following the ideas of the above proof we can set some reference point a in S and define a function Ue S R byU e r g a r F e u d u displaystyle U e mathbf r int gamma mathbf a mathbf r mathbf F e mathbf u cdot mathrm d mathbf u nbsp Using the above proof we know Ue is well defined and differentiable and Fe Ue from this formula we can use the gradient theorem to easily derive the well known formula for calculating work done by conservative forces W DU This function Ue is often referred to as the electrostatic potential energy of the system of charges in S with reference to the zero of potential a In many cases the domain S is assumed to be unbounded and the reference point a is taken to be infinity which can be made rigorous using limiting techniques This function Ue is an indispensable tool used in the analysis of many physical systems Generalizations editMain articles Stokes theorem and Closed and exact differential forms Many of the critical theorems of vector calculus generalize elegantly to statements about the integration of differential forms on manifolds In the language of differential forms and exterior derivatives the gradient theorem states that g ϕ g d ϕ displaystyle int partial gamma phi int gamma mathrm d phi nbsp for any 0 form ϕ defined on some differentiable curve g Rn here the integral of ϕ over the boundary of the g is understood to be the evaluation of ϕ at the endpoints of g Notice the striking similarity between this statement and the generalized Stokes theorem which says that the integral of any compactly supported differential form w over the boundary of some orientable manifold W is equal to the integral of its exterior derivative dw over the whole of W i e W w W d w displaystyle int partial Omega omega int Omega mathrm d omega nbsp This powerful statement is a generalization of the gradient theorem from 1 forms defined on one dimensional manifolds to differential forms defined on manifolds of arbitrary dimension The converse statement of the gradient theorem also has a powerful generalization in terms of differential forms on manifolds In particular suppose w is a form defined on a contractible domain and the integral of w over any closed manifold is zero Then there exists a form ps such that w dps Thus on a contractible domain every closed form is exact This result is summarized by the Poincare lemma See also editState function Scalar potential Jordan curve theorem Differential of a function Classical mechanics Line integral Path independence Conservative vector field Path independenceReferences edit Williamson Richard and Trotter Hale 2004 Multivariable Mathematics Fourth Edition p 374 Pearson Education Inc Stewart James 2015 16 3 The Fundamental Theorem for Line Integrals Calculus 8th ed Cengage Learning pp 1127 1128 ISBN 978 1 285 74062 1 a b Williamson Richard and Trotter Hale 2004 Multivariable Mathematics Fourth Edition p 410 Pearson Education Inc Retrieved from https en wikipedia org w index php title Gradient theorem amp oldid 1142875125, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.