Skip to main content

Linear Elliptic Equations

  • Chapter
  • First Online:
Partial Differential Equations I

Part of the book series: Applied Mathematical Sciences ((AMS,volume 115))

  • 11k Accesses

Abstract

The first major topic of this chapter is the Dirichlet problem for the Laplace operator on a compact domain with boundary:

$$\Delta u = 0\text{ on }\Omega,\quad {u\bigr |}_{\partial\Omega } = f.$$
(0.1)

We also consider the nonhomogeneous problem Δu = g and allow for lower-order terms. As in Chap. 2, Δ is the Laplace operator determined by a Riemannian metric. In §1 we establish some basic results on existence and regularity of solutions, using the theory of Sobolev spaces. In §2 we establish maximum principles, which are useful for uniqueness theorems and for treating (0.1) for f continuous, among other things.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. R. Adams, Sobolev Spaces, Academic, New York, 1975.

    MATH  Google Scholar 

  2. S. Agmon, Lectures on Elliptic Boundary Problems, Van Nostrand, New York, 1964.

    Google Scholar 

  3. S. Agmon, A. Douglis, and L. Nirenberg, Estimates near the boundary for solutions of elliptic differential equations satisfying general boundary conditionås, CPAM 12(1959), 623–727; II, CPAM 17(1964), 35–92.

    Google Scholar 

  4. L. Ahlfors, Complex Analysis, McGraw-Hill, New York, 1979.

    MATH  Google Scholar 

  5. T. Aubin, Nonlinear Analysis on Manifolds. Monge-Ampere Equations, Springer, New York, 1982.

    Google Scholar 

  6. H. Bateman, Partial Differential Equations of Mathematical Physics, Dover, New York, 1944.

    MATH  Google Scholar 

  7. M. Berger, P. Gauduchon, and E. Mazet, Le Spectre d’une Variété Riemannienne, LNM no. 194, Springer, New York, 1971.

    Google Scholar 

  8. L. Bers, F. John, and M. Schechter, Partial Differential Equations, Wiley, New York, 1964.

    MATH  Google Scholar 

  9. R. Bott and L. Tu, Differential Forms in Algebraic Topology, Springer, New York, 1982.

    Book  Google Scholar 

  10. F. Browder, A priori estimates for elliptic and parabolic equations, AMS Proc. Symp. Pure Math. IV(1961), 73–81.

    Google Scholar 

  11. H. O. Cordes, Elliptic Pseudodifferential Operators – an Abstract Theory, LNM no. 756, Springer, New York, 1979.

    Google Scholar 

  12. R. Courant and D. Hilbert, Methods of Mathematical Physics II, Wiley, New York, 1966.

    MATH  Google Scholar 

  13. B. Dahlberg and C. Kenig, Hardy spaces and the Neumann problem in Lp for Laplace’s equation in Lipschitz domains, Ann. Math. 125(1987), 437–465.

    Google Scholar 

  14. G. Folland, Introduction to Partial Differential Equations, Princeton University Press, Princeton, N. J., 1976.

    MATH  Google Scholar 

  15. G. Folland and J. J. Kohn, The Neumann Problem for the Cauchy-Riemann Complex, Princeton University Press, Princeton, N. J., 1972.

    MATH  Google Scholar 

  16. A. Friedman, Generalized Functions and Partial Differential Equations, Prentice-Hall, Englewood Cliffs, N. J., 1963.

    Book  Google Scholar 

  17. K. Friedrichs, On the differentiability of the solutions of linear elliptic equations, CPAM 6(1953), 299–326.

    Google Scholar 

  18. P. Garabedian, Partial Differential Equations, Wiley, New York, 1964.

    MATH  Google Scholar 

  19. L. Gårding, Dirichlet’s problem for linear elliptic partial differential equations, Math. Scand. 1(1953), 55–72.

    Google Scholar 

  20. D. Gilbarg and N. Trudinger, Elliptic Partial Differential Equations of Second Order, 2nd ed., Springer, New York, 1983.

    Book  Google Scholar 

  21. P. Grisvard, Elliptic Problems in Nonsmooth Domains, Pitman, Boston, 1985.

    MATH  Google Scholar 

  22. L. Helms, Introduction to Potential Theory, Wiley, New York, 1969.

    MATH  Google Scholar 

  23. E. Hille, Analytic Function Theory, Chelsea, New York, 1977.

    Google Scholar 

  24. W. Hodge, The Theory and Applications of Harmonic Integrals, Cambridge University Press, Cambridge, 1952.

    MATH  Google Scholar 

  25. L. Hörmander, The Analysis of Linear Partial Differential Operators, Vols. 3 and 4, Springer, New York, 1985.

    Google Scholar 

  26. L. Hörmander, Introduction to Complex Analysis in Several Variables, Van Nostrand, Princeton, N. J. 1966.

    MATH  Google Scholar 

  27. F. John, Partial Differential Equations, Springer, New York, 1975.

    Book  Google Scholar 

  28. O. Kellogg, Foundations of Potential Theory, Dover, New York, 1954.

    MATH  Google Scholar 

  29. S. Kichenassamy, Compactness theorems for differential forms, CPAM 42(1989), 47–53.

    Google Scholar 

  30. O. Ladyzhenskaya and N. Ural’tseva, Linear and Quasilinear Elliptic Equations, Academic, New York, 1968.

    MATH  Google Scholar 

  31. H. Lebesgue, Sur le problème de Dirichlet, Rend. Circ. Mat. Palermo 24(1907), 371–402.

    Google Scholar 

  32. J. Lions and E. Magenes, Non-homogeneous Boundary Problems and Applications I, II, Springer, New York, 1972.

    Book  Google Scholar 

  33. J. Milnor, Topology from the Differentiable Viewpoint, University Press of Virginia, Charlottesville, VA, 1965.

    MATH  Google Scholar 

  34. C. Miranda, Partial Differential Equations of Elliptic Type, Springer, New York, 1970.

    Book  Google Scholar 

  35. S. Mizohata, The Theory of Partial Differential Equations, Cambridge University Press, Cambridge, 1973.

    MATH  Google Scholar 

  36. C. B. Morrey, Multiple Integrals in the Calculus of Variations, Springer, New York, 1966.

    MATH  Google Scholar 

  37. F. Murat, Compacité par compensation, Ann. Scuola Norm. Sup. Pisa 5(1978), 485–507.

    Google Scholar 

  38. L. Nirenberg, On elliptic partial differential equations, Ann. Scuola Norm. Sup. Pisa 13(1959), 116–162.

    Google Scholar 

  39. L. Nirenberg, Estimates and existence of solutions of elliptic equations, CPAM 9(1956), 509–530.

    Google Scholar 

  40. L. Nirenberg, Lectures on Linear Partial Differential Equations, Reg. Conf. Ser. Math., no. 17, AMS, Providence, R.I., 1972.

    Google Scholar 

  41. M. Protter and H. Weinberger, Maximum Principles in Differential Equations, Springer, New York, 1984.

    Book  Google Scholar 

  42. J. Rauch, Partial Differential Equations, Springer, New York, 1992.

    MATH  Google Scholar 

  43. J. Rauch and M. Taylor, Potential and scattering theory on wildly perturbed domains, J. Funct. Anal.. 18(1975), 27–59.

    Google Scholar 

  44. M. Reed and B. Simon, Methods of Mathematical Physics, Academic, New York, Vols. 1,2, 1975; Vols. 3,4, 1978.

    Google Scholar 

  45. J. Robbin, R. Rogers, and B. Temple, On weak continuity and the Hodge decomposition, Trans. AMS 303(1987), 609–618.

    Google Scholar 

  46. W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 1976.

    MATH  Google Scholar 

  47. M. Schechter, Modern Methods in Partial Differential Equations, an Introduction, McGraw-Hill, New York, 1977.

    MATH  Google Scholar 

  48. J. Serrin, The problem of Dirichlet for quasilinear elliptic differential equations with many independent variables, Phil. Trans. R. Soc. Lond. Ser. A 264(1969), 413–496.

    Google Scholar 

  49. I. M. Singer and J. Thorpe, Lecture Notes on Elementary Topology and Geometry, Springer, New York, 1976.

    Book  Google Scholar 

  50. S. Sobolev, Partial Differential Equations of Mathematical Physics, Dover, New York, 1964.

    MATH  Google Scholar 

  51. I. Stakgold, Boundary Value Problems of Mathematical Physics, Macmillan, New York, 1968.

    MATH  Google Scholar 

  52. J. J. Stoker, Differential Geometry, Wiley-Interscience, New York, 1969.

    MATH  Google Scholar 

  53. L. Tartar, Compensated compactness and applications to partial differential equations, in Nonlinear Analysis and Mechanics (R.Knops, ed.), Research Notes in Math., Pitman, Boston, 1979.

    Google Scholar 

  54. F. Treves, Basic Linear Partial Differential Equations, Academic, New York, 1975.

    MATH  Google Scholar 

  55. M. Tsuji, Potential Theory, Chelsea, New York, 1975.

    MATH  Google Scholar 

  56. F. Warner, Foundations of Differentiable Manifolds and Lie Groups, Scott, Foresman, Glenview, Ill., 1971.

    Google Scholar 

  57. C. Wilcox, Scattering Theory for the d’Alembert Equation in Exterior Domains, LNM no. 442, Springer, New York, 1975.

    Google Scholar 

  58. K. Yosida, Functional Analysis, Springer, New York, 1965.

    Book  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael E. Taylor .

Appendices

A Spaces of generalized functions on manifolds with boundary

Let \(\overline{M}\) be a compact manifold with smooth boundary. We will define a one-parameter family of spaces of functions and “generalized functions” on M, analogous to the Sobolev spaces defined when \(\partial M =\varnothing \). The spaces will be defined in terms of a Laplace operator Δ on M, and a boundary condition for the Laplace operator. We will explicitly discuss only the Dirichlet boundary condition, though the results given work equally well for other coercive boundary conditions yielding self-adjoint operators, such as the Neumann boundary condition.

Fixing on the Dirichlet boundary condition, let us recall from (1.7) the map

$$T : {H}^{-1}(M)\rightarrow {H}^{1}(M), $$
(A.1)

inverting the Laplace operator

$$\Delta : {H}_{0}^{1}(M)\rightarrow {H}^{-1}(M). $$
(A.2)

The restriction of T to L2(M) is compact and self-adjoint, and we have an orthonormal basis of L2(M) consisting of eigenfunctions:

$${u}_{j}\in{H}_{0}^{1}(M)\cap{C}^{\infty }(\overline{M}),\quad T{u}_{ j} = -{\mu }_{j}{u}_{j},\quad\Delta {u}_{j} = -{\lambda }_{j}{u}_{j}, $$
(A.3)

where μj ↘ 0, 0 < λj.

For a given vL2(M), set

$$v =\sum\limits_{j}\hat{v}(j)\ {u}_{j},\quad\hat{v}(j) = (v,{u}_{j}). $$
(A.4)

Now, for s ≥ 0, we define

$$\begin{array}{rcl}{\mathcal{D}}_{s }& =& {\Bigl\{v\in{L}^{2}(M) :\sum\limits_{j\geq 0}\vert\hat{v}(j){\vert }^{2}{\lambda }_{ j}^{s} <\infty\Bigr\}}\\ & =& {\Bigl\{v\in{L}^{2}(M) :\sum\limits_{j\geq 0}\hat{v}(j){\lambda }_{j}^{s/2}{u}_{ j}\in{L}^{2}(M)\Bigr\}}.\end{array}$$
(A.5)

In view of (A.3), an equivalent characterization is

$${\mathcal{D}}_{s } = {(-\mathcal{T} )}^{s /2}{\mathcal{L}}^{2}(M). $$
(A.6)

Clearly, we have

$${\mathcal{D}}_{0} = {\mathcal{L}}^{2}(M). $$
(A.7)

Also, \({\mathcal{D}}_{2} =\mathcal{T}\ {\mathcal{L}}^{2}(M)\), and by Theorem 1.3 we have

$${\mathcal{D}}_{2} = {\mathcal{H}}^{2}(M)\cap {\mathcal{H}}_{0}^{1}(M). $$
(A.8)

Generally, \({\mathcal{D}}_{s + 2 } =\mathcal{T}\ {\mathcal{D}}_{s }\), so Theorem 1.3 also gives, inductively,

$$\mathcal{D}_{2k} \subset H^{2k} (M), k = 1, 2, 3, \ldots .$$
(A.9)

A result perhaps slightly less obvious than (A.7)–(A.9) is that

$${\mathcal{D}}_{1 } = {\mathcal{H}}_{0}^{\infty }(M).$$
(A.10)

To see this, note that \({\mathcal{D}}_{s }\) is the completion of the space \(\mathcal{F}\) of finite linear combinations of the eigenfunctions {uj}, with respect to the \({\mathcal{D}}_{s }\)-norm, defined by

$$\|{v\|}_{{\mathcal{D}}_{s }}^{2} =\sum\limits_{j}\vert\hat{v}(j){\vert }^{2}{\lambda }_{ j}^{s}. $$
(A.11)

Now, if \(v\in\mathcal{F}\), then

$$(dv,dv) = (v,-\Delta v) =\sum \nolimits (v,{u}_{j})({u}_{j},-\Delta v) =\sum \nolimits\vert\hat{v}(j){\vert }^{2}{\lambda }_{ j}, $$
(A.12)

so

$$\|{v\|}_{{\mathcal{D}}_{1 }}^{2} =\| d{v\|}_{{ L}^{2}(M)}^{2}, $$
(A.13)

for \(v\in\mathcal{F}\). In fact, \({\mathcal{D}}_{s }\) is the completion of \({\mathcal{D}}_{\sigma }\) in the \({\mathcal{D}}_{s }\)-norm for any σ > s. We see that (A.13) holds for all \(v\in {\mathcal{D}}_{2}\), and, with \({\mathcal{D}}_{2}\) characterized by (A.8), it is clear that the completion in the norm (A.13) is described by (A.10).

If the Neumann boundary condition were considered, we would replace λj by ⟨λj⟩ to take care of λ0 = 0. In such a case, we would have

$${\mathcal{D}}_{2} ={\bigl\{u \in {\mathcal{H}}^{2}(M) :\frac{\partial u } {\partial\nu } = 0\text{ on }\partial M \Bigr\}},\quad {\mathcal{D}}_{1 } = {\mathcal{H}}^{1}(M).$$

Now, for s < 0, we define \({\mathcal{D}}_{s }\) to be the dual of \({\mathcal{D}}_{-s}\):

$${\mathcal{D}}_{s } = {\mathcal{D}}_{-s}^{{\ast}}. $$
(A.14)

In particular, for any \(v\in {\mathcal{D}}_{s }\), and any \(s\in \mathbb{R}\), \((v,{u}_{j}) =\hat{ v}(j)\) is defined, and we see that the characterizations involving the sums in (A.5) continue to hold for all \(s\in \mathbb{R}\). Also the norm (A.11) provides a Hilbert space structure on \({\mathcal{D}}_{s }\) for all \(s\in \mathbb{R}\). By (A.10) we have (for Dirichlet boundary conditions)

$${\mathcal{D}}_{-1} = {\mathcal{H}}^{-1 }(M). $$
(A.15)

Also, we have the interpolation identity

$${[{\mathcal{D}}_{s },{\mathcal{D}}_{\sigma }]}_{\theta } = {\mathcal{D}}_{\theta\sigma + (1 -\theta )s}, $$
(A.16)

for all \(s,\sigma \in \mathbb{R},\theta \in [0,1]\), where the interpolation spaces are as defined in Chap. 4.

The isomorphism

$$\Delta : {\mathcal{D}}_{s + 2}\rightarrow {\mathcal{D}}_{s },\text{ with inverse }\mathcal{T} : {\mathcal{D}}_{s }\rightarrow {\mathcal{D}}_{s + 2}, $$
(A.17)

obviously valid for s ≥ 0, extends by duality to an isomorphism \(\Delta : {\mathcal{D}}_{-s}\rightarrow {\mathcal{D}}_{- s -2 }\) for s ≥ 0, so (A.17) also holds for s ≤ − 2. By interpolation, it holds for all real s.

By interpolation, (A.9) implies

$${\mathcal{D}}_{s }\subset {\mathcal{H}}^{s }(M),\text{ for } s\ \geq 0. $$
(A.18)

The natural map \({\mathcal{D}}_{s }\hookrightarrow {\mathcal{H}}^{s}(M)\) is injective, for s ≥ 0, but it is not generally onto, and the transpose \({H}^{-s}(M)\rightarrow {\mathcal{D}}_{-s}\) is not generally injective. However, the natural map

$${H}_{\text{ comp}}^{-s}(M)\rightarrow {\mathcal{D}}_{ -s} $$
(A.19)

is injective, where \({H}_{\text{ comp}}^{-s}(M)\) denotes the space of elements of Hs(N) (N being the double of M) with support in the interior of M. In particular, for any interior point pM,

$${\delta }_{p}\in {\mathcal{D}}_{-s},\text{ for } s > \frac{n}{2}\quad (n =\text{ dim } M ). $$
(A.20)

Note that as p∂M, δp → 0 in any of these spaces. From the isomorphism in (A.17), we have

$${G}_{p} = {\Delta }^{-1}{\delta }_{ p} = T{\delta }_{p} $$
(A.21)

well defined, and

$${G}_{p}\in {\mathcal{D}}_{-n /2 +2 -\epsilon },\text{ for all }\epsilon > 0. $$
(A.22)

This object is equivalent to the Green function studied in this chapter.

We can write any \(v\in {\mathcal{D}}_{s }\), even for s < 0, as a Fourier series with respect to the eigenfunctions uj. In fact, defining \(\hat{v}(j) = (v,{u}_{j})\), as before, the series \({\sum}_{j}\hat{v}(j){u}_{j}\) is convergent in the space \({\mathcal{D}}_{s }\) to v, provided \(v\in {\mathcal{D}}_{s }\), so we are justified in writing

$$v =\sum\limits_{j}\hat{v}(j){u}_{j},\quad v\in {\mathcal{D}}_{s },\text{ for any } s \in \mathbb{R}. $$
(A.23)

Note that \(-\Delta : {\mathcal{D}}_{s }\rightarrow {\mathcal{D}}_{s -\in }\) is given by

$$-\Delta v =\sum\limits_{j}{\lambda }_{j}\hat{v}(j){u}_{j}, $$
(A.24)

for any \(s\in \mathbb{R}\). We can define

$${(-\Delta )}^{(\sigma +i\tau )} : {\mathcal{D}}_{s }\rightarrow {\mathcal{D}}_{s -2 \sigma }, $$
(A.25)

for any \(\sigma,\tau \in \mathbb{R}\), by

$${(-\Delta )}^{(\sigma +i\tau )}v =\sum \nolimits {\lambda }_{j}^{(\sigma +i\tau )}\hat{v}(j){u}_{ j}, $$
(A.26)

where \(v\in {\mathcal{D}}_{s }\) is given by (A.23). The maps (A.25) are all isomorphisms. Note that we can write the \({\mathcal{D}}_{s }\)-inner product coming from (A.11) as

$${(v,w)}_{{\mathcal{D}}_{s }} ={\bigl ( v,{(-\Delta )}^{s}w\bigr )}, $$
(A.27)

where on the right side the pairing arises from the natural \({\mathcal{D}}_{s } : {\mathcal{D}}_{-s}\) duality.

B The Mayer–Vietoris sequence in deRham cohomology

Here we establish a useful complement to the long exact sequence (9.67) and illustrate some of its implications. Let X be a smooth manifold, and suppose X is the union of two open sets, M1 and M2. Let U = M1M2. The Mayer–Vietoris sequence has the form

$$\cdots \rightarrow {\mathcal{H}}^{k-1}(\overline{U})\mathop{\longrightarrow}\limits_{}^{\delta }{\mathcal{H}}^{k}(X)\mathop{\longrightarrow}\limits_{}^{\rho }{\mathcal{H}}^{k}({\overline{M}}_{ 1}) \oplus {\mathcal{H}}^{k}({\overline{M}}_{ 2})\mathop{\longrightarrow}\limits_{}^{\gamma }{\mathcal{H}}^{k}(\overline{U}) \rightarrow \cdots.$$
(B.1)

These maps are defined as follows. A closed form α ∈ Λk(X) restricts to a pair of closed forms on M1 and M2, yielding ρ in a natural fashion. The map γ also comes from restriction; if ιν : UMν, a pair of closed forms αν ∈ Λk(Mν) goes to \({\iota }_{1}^{{\ast}}{\alpha }_{1} - {\iota }_{2}^{{\ast}}{\alpha }_{2}\), defining γ. Clearly, \({\iota }_{1}^{{\ast}}(\alpha {\vert }_{{M}_{1}}) = {\iota }_{2}^{{\ast}}(\alpha {\vert }_{{M}_{2}})\) if α ∈ Λk(X), so γ ∘ ρ = 0.

To define the “coboundary map” δ on a class [α], with α ∈ Λk(U) closed, pick βν ∈ Λk(Mν) such that \(\alpha = {\beta }_{1} - {\beta }_{2}\). Thus dβ1 = dβ2 on U. Set

$$\delta [\alpha ] = [\sigma ]\text{ with }\sigma = d{\beta }_{\nu }\text{ on }{M}_{\nu }. $$
(B.2)

To show that (B.2) is well defined, suppose βν ∈ Λk(Mν) and \({\beta }_{1} - {\beta }_{2} = d\omega \) on U. Let {φν} be a smooth partition of unity supported on {Mν}, and consider \(\psi = {\varphi }_{1}{\beta }_{1} + {\varphi }_{2}{\beta }_{2}\), where φνβν is extended by 0 off Mν. We have \(d\psi = {\varphi }_{1}d{\beta }_{1} + {\varphi }_{2}d{\beta }_{2} + d{\varphi }_{1}\wedge ({\beta }_{1} - {\beta }_{2}) =\sigma + d{\varphi }_{1}\wedge ({\beta }_{1} - {\beta }_{2}).\) Since dφ1 is supported on U, we can write

$$\sigma = d\psi - d(d{\varphi }_{1}\wedge \omega ),$$

an exact form on X, so (B.2) makes δ well defined. Obviously, the restriction of σ to each Mν is always exact, so ρ ∘ δ = 0. Also, if \(\alpha = {\iota }_{1}^{{\ast}}{\alpha }_{1} - {\iota }_{2}^{{\ast}}{\alpha }_{2}\) on U, we can pick βν = αν to define δ[α]. Then \(d{\beta }_{\nu } = d{\alpha }_{\nu } = 0\), so δ ∘ γ = 0.

In fact, the sequence (B.1) is exact, that is,

$$\text{ im }\delta =\text{ ker }\rho,\quad\text{ im }\rho =\text{ ker }\gamma,\quad\text{ im }\gamma =\text{ ker }\delta. $$
(B.3)

We leave the verification of this as an exercise, which can be done with arguments similar to those sketched in Exercises 11–13 in the exercises on cohomology after §9.

If Mν are the interiors of compact manifolds with smooth boundary, and \(\overline{U} =\overline{{M}_{1}}\cap\overline{{M}_{2}}\) has smooth boundary, the argument above extends directly to produce an exact sequence

$$\cdots \rightarrow \mathcal{H}^{k-1} (\overline{U}) \mathop{\rightarrow}\limits^{\delta}\mathcal{H}^{k} (X) \mathop{\rightarrow}\limits^{\rho} \mathcal{H}^{k} (\overline{M}_1) \oplus \mathcal{H}^{k} (\overline{M}_2) \mathop{\rightarrow}\limits^{\gamma} \mathcal{H}^{k} (\overline{U}) \rightarrow \cdots .$$
(B.4)

Furthermore, suppose that instead \(X ={\overline{M}}_{1}\cup {\overline{M}}_{2}\) and \({\overline{M}}_{1}\cap {\overline{M}}_{2} = Y\) is a smooth hypersurface in X. One also has an exact sequence

$$\cdots \rightarrow \mathcal{H}^{k-1} (\overline{Y}) \mathop{\rightarrow}\limits^{\delta}\mathcal{H}^{k} (X) \mathop{\rightarrow}\limits^{\rho} \mathcal{H}^{k} (\overline{M}_1) \oplus \mathcal{H}^{k} (\overline{M}_2) \mathop{\rightarrow}\limits^{\gamma} \mathcal{H}^{k} (Y) \rightarrow \cdots .$$
(B.5)

To relate (B.4) and (B.5), let U be a collar neighborhood of Y, and form (B.4) with \({\overline{M}}_{\nu }\) replaced by \({\overline{M}}_{\nu }\cup\overline{U}\). There is a map \(\pi :\overline{U}\rightarrow Y\), collapsing orbits of a vector field transversal to Y, and π induces an isomorphism of cohomology groups, \({\pi }^{{\ast}} : {\mathcal{H}}^{k }(\overline{U})\approx {\mathcal{H}}^{k }(\mathcal{Y})\).

To illustrate the use of (B.5), suppose \(X = {S}^{n},Y = {S}^{n-1}\) is the equator, and \({\overline{M}}_{\nu }\) are the upper and lower hemispheres, each diffeomorphic to the ball \(\overline{{B}^{n}}\). Then we have an exact sequence

$$\begin{array}{l} \cdots \rightarrow \mathcal{H}^{k-1} (\overline{B^n}) \oplus \mathcal{H}^{k-1} (\overline{B^n}) \mathop{\rightarrow}\limits^{\gamma} \mathcal{H}^{k-1} (S^{n-1}) \mathop{\rightarrow}\limits^{\delta} \mathcal{H}^{k} (S^n) \\ \qquad \qquad \qquad \qquad \qquad \qquad \qquad \mathop{\rightarrow}\limits^{\rho} \mathcal{H}^k (\overline{B^n}) \oplus \mathcal{H}^k (\overline{B^n}) \rightarrow \cdots . \end{array}$$
(B.6)

As in (9.70), \({\mathcal{H}}^{k }(\overline{{B}^{n}}) = 0\) except for k = 0, when you get \(\mathbb{R}\). Thus

$$\delta : {\mathcal{H}}^{k - 1}({\mathcal{S}}^{n - 1})\mathop{\longrightarrow}\limits_{}^{\approx }{\mathcal{H}}^{k}({\mathcal{S}}^{n}),\text{ for } k > 1. $$
(B.7)

Granted that the computation \({\mathcal{H}}^{1}({\mathcal{S}}^{1})\approx \mathbb{R}\) is elementary, this implies \({\mathcal{H}}^{n }({\mathcal{S}}^{n })\approx \mathbb{R}\), for n ≥ 1. Looking at the segment

$$0\rightarrow {\mathcal{H}}^{0}({\mathcal{S}}^{n })\mathop{\longrightarrow}\limits_{}^{\rho }{\mathcal{H}}^{0}(\overline{{B}^{n}})\oplus {\mathcal{H}}^{0}(\overline{{B}^{n}})\mathop{\longrightarrow}\limits_{}^{\gamma }{\mathcal{H}}^{0}({\mathcal{S}}^{n - 1 })\mathop{\longrightarrow}\limits_{}^{\delta }{\mathcal{H}}^{1}({\mathcal{S}}^{n })\rightarrow 0,$$

we see that if n ≥ 2, then ker \(\gamma \approx \mathbb{R}\), so γ is surjective, hence δ = 0, so \({\mathcal{H}}^{1}({\mathcal{S}}^{n }) = 0\), for n ≥ 2. Also, if 0 < k < n, we see by iterating (B.7) that \({\mathcal{H}}^{k}({\mathcal{S}}^{n})\approx {\mathcal{H}}^{1}({\mathcal{S}}^{n - k + 1})\), so \({\mathcal{H}}^{k}({\mathcal{S}}^{n}) = 0\), for 0 < k < n. Since obviously \({\mathcal{H}}^{0}({\mathcal{S}}^{n }) =\mathbb{R}\) for n ≥ 1, we have a fourth computation of \({\mathcal{H}}^{k}({\mathcal{S}}^{n})\), distinct from those sketched in Exercise 10 of §8 and in Exercises 10 and 14 of the set of exercises on cohomology after §9.

We note an application of (B.5) to the computation of Euler characteristics, namely

$$\chi ({\overline{M}}_{1}) +\chi ({\overline{M}}_{2}) =\chi (X) +\chi (Y ). $$
(B.8)

Note that this result contains some of the implications of Exercises 17 and 18 in the exercises on cohomology, in §9.

Using this, it is an exercise to show that if one two-dimensional surface X1 is obtained from another X0 by adding a handle, then \(\chi ({X}_{1}) =\chi ({X}_{0}) - 2\). In particular, if Mg is obtained from S2 by adding g handles, then \(\chi ({M}^{g}) = 2 - 2g\). Thus, if Mg is orientable, since \({\mathcal{H}}^{0}({ M }^{g})\approx {\mathcal{H}}^{2}({ M }^{g})\approx \mathbb{R}\), we have

$${\mathcal{H}}^{1}({ M }^{g})\approx{\mathbb{R}}^{2g}. $$
(B.9)

It is useful to examine the beginning of the sequence (B.5):

$$0 \rightarrow \mathcal{H}^0 (X) \mathop{\rightarrow}\limits^{\rho} \mathcal{H}^0 (\overline{M}_1) \oplus \mathcal{H}^0 (\overline{M}_2) \mathop{\rightarrow}\limits^{\gamma} \mathcal{H}^0 (Y) \mathop{\rightarrow}\limits^{\delta} \mathcal{H}^1 (X) \rightarrow \cdots ,$$
(B.10)

Suppose C is a smooth, closed curve in S2. Apply (B.10) with \({M}_{1} =\mathcal{C}\), a collar neighborhood of C, and \({\overline{M}}_{2} =\overline{\Omega }\), the complement of \(\mathcal{C}\). Since \(\partial\mathcal{C}\) is diffeomorphic to two copies of C, and since \({\mathcal{H}}^{1}({\mathcal{S}}^{2}) = 0\), (B.10) becomes

$$0\rightarrow \mathbb{R}\mathop{\longrightarrow}\limits_{}^{\rho }\mathbb{R}\oplus {\mathcal{H}}^{0}(\overline{\Omega })\mathop{\longrightarrow}\limits_{}^{\gamma }\mathbb{R}\oplus \mathbb{R}\mathop{\longrightarrow}\limits_{}^{\delta } 0. $$
(B.11)

Thus γ is surjective while ker γ = im \(\rho \approx \mathbb{R}\). This forces

$${\mathcal{H}}^{0}(\overline{\Omega })\approx \mathbb{R}\oplus \mathbb{R}. $$
(B.12)

In other words, \(\overline{\Omega }\) has exactly two connected components. This is the smooth case of the Jordan curve theorem. Jordan’s theorem holds when C is a homeomorphic image of S1, but the trick of putting a collar about C does not extend to this case.

More generally, if X is a compact, connected, smooth, oriented manifold such that \({\mathcal{H}}^{1}(\mathcal{X}) = 0\), and if Y is a smooth, compact, connected, oriented hypersurface, then letting \(\mathcal{C}\) be a collar neighborhood of Y and \(\overline{\Omega } = X\setminus\mathcal{C}\), we again obtain the sequence (B.11) and hence the conclusion (B.12). The orientability ensures that \(\partial\mathcal{C}\) is diffeomorphic to two copies of Y. This produces the following variant of (the smooth case of) the Jordan–Brouwer separation theorem.

Theorem B.1.

If X is a smooth manifold, Y is a smooth submanifold of codimension 1, both are

$$compact,\ connected,\ and\ oriented,$$

and

$${\mathcal{H}}^{1}(\mathcal{X}) = 0,$$

then X ∖ Y has precisely two connected components.

If all these conditions hold, except that Y is not orientable, then we replace \(\mathbb{R}\oplus \mathbb{R}\) by \(\mathbb{R}\) in (B.11) and conclude that XY is connected, in that case. As an example, the real projective space \(\mathbb{R}{\mathbb{P}}^{2}\) sits in \(\mathbb{R}{\mathbb{P}}^{3}\) in such a fashion.

Recall from §19 of Chap. 1 the elementary proof of Theorem B.1 when \(X = {\mathbb{R}}^{n+1}\), in particular the argument using degree theory that if Y is a compact, oriented surface in \({\mathbb{R}}^{n+1}\) (hence, in Sn + 1), then its complement has at least two connected components. One can extend the degree-theory argument to the nonorientable case, as follows.

There is a notion of degree mod 2 of a map F: YSn, which is well defined whether or not Y is orientable. For one approach, see [Mil]. This is also invariant under homotopy. Now, if in the proof of Theorem 19.11 of Chap. 1, one drops the hypothesis that the hypersurface Y (denoted X there) is orientable, it still follows that the mod 2 degree of Fp must jump by±1 when p crosses Y, so \({\mathbb{R}}^{n+1}\setminus Y\) still must have at least two connected components. In view of the result noted after Theorem B.1, this situation cannot arise. This establishes the following.

Proposition B.2.

If Y is a compact hypersurface of \({\mathbb{R}}^{n+1}\) (or S n+1 ), then Y is orientable.

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Taylor, M.E. (2011). Linear Elliptic Equations. In: Partial Differential Equations I. Applied Mathematical Sciences, vol 115. Springer, New York, NY. https://doi.org/10.1007/978-1-4419-7055-8_5

Download citation

Publish with us

Policies and ethics