Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


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 References  














Gradient theorem






العربية
Bosanski
Español
Esperanto
فارسی
Français

Italiano
עברית
Português
Українська


 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


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.

Ifφ : URnR is a differentiable function and γadifferentiable 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]

Ifφ is a differentiable function from some open subset URntoR and r is a differentiable function from some closed interval [a, b]toU (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 tin[a, b]. Here the denotes the usual inner product.

Now suppose the domain Uofφ contains the differentiable curve γ with endpoints p and q. (This is oriented in the direction from ptoq). If r parametrizes γ for tin[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 ptoq. For uinRn, let |u| denote the Euclidean normofu. 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 Rnifα ≥ 1.

Ifα <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 piinR3. We would like to calculate the work done on a particle of charge q as it travels from a point a to a point binR3. 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 uinR3, and k = 1/(4πε0), where ε0 is the vacuum permittivity.

Let γR3 − {p1, ..., pn} be an arbitrary differentiable curve from atob. 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 potentialorelectrostatic 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 Fisconservative), 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 — IfF 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 aofU, and define f : URbyHere γ[a, x] is any (differentiable) curve in U originating at a and terminating at x. We know that fiswell-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 becomeswhere 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 viswhere 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 ainS, and define a function Ue: SRby

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 formsonmanifolds. 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.
  • ^ Stewart, James (2021). "16.3 The Fundamental Theorem for Line Integrals". Calculus (9th ed.). Cengage Learning. pp. 1182–1185. ISBN 978-1-337-62418-3.
  • ^ 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&oldid=1217081265"

    Category: 
    Theorems in calculus
    Hidden categories: 
    Articles with short description
    Short description is different from Wikidata
    Pages using sidebar with the child parameter
    Articles containing proofs
     



    This page was last edited on 3 April 2024, at 18:43 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki