Skip to main content
Log in

Heat flow, heat content and the isoparametric property

  • Published:
Mathematische Annalen Aims and scope Submit manuscript

Abstract

Let M be a Riemannian manifold and \(\Omega \) a compact domain of M with smooth boundary. We study the solution of the heat equation on \(\Omega \) having constant unit initial conditions and Dirichlet boundary conditions. The purpose of this paper is to study the geometry of domains for which, at any fixed value of time, the normal derivative of the solution (heat flow) is a constant function on the boundary. We express this fact by saying that such domains have the constant flow property. In constant curvature spaces known examples of such domains are given by geodesic balls and, more generally, by domains whose boundary is connected and isoparametric. The question is: are they all like that? This problem is the analogous (for the heat equation) of the classical Serrin’s problem for harmonic domains. In this paper we give an affirmative answer to the above question: in fact we prove more generally that, if a domain in an analytic Riemannian manifold has the constant flow property, then every component of its boundary is an isoparametric hypersurface. For space forms, we also relate the order of vanishing of the heat content with fixed boundary data with the constancy of the r-mean curvatures of the boundary and with the isoparametric property. Finally, we discuss the constant flow property in relation to other well-known overdetermined problems involving the Laplace operator, like the Serrin problem or the Schiffer problem.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Similar content being viewed by others

References

  1. Berenstein, C.A.: An inverse spectral theorem and its applications to the Pompeiu problem. J. Anal. Math. 37, 128–144 (1980)

    Article  MathSciNet  MATH  Google Scholar 

  2. Berenstein, C.A.: On an overdetermined Neumann problem. Geom Semin. Univ. Stud, Bologna, pp 21–26 (1986)

  3. Berenstein, C.A., Shahshahani, M.: Harmonic analysis and the Pompeiu problem. Am. J. Math. 105, 1217–1229 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  4. Berenstein, C., Yang, P.: An inverse Neumann problem. J. Reine Angew. Math. 382, 1–21 (1987)

    MathSciNet  MATH  Google Scholar 

  5. Caffarelli, L.: The regularity of free boundaries in higher dimensions. Acta Math. 139, 155–184 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  6. Capparelli, S., Maroscia, P.: On two sequences of orthogonal polynomials related to Jordan blocks. Mediterr. J. Math. 10(4), 1609–1630 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  7. Cartan, E.: Familles de surfaces isoparamétriques dans les espaces à courbure constante. Annali di Mat. 17, 177–191 (1938)

    Article  MathSciNet  MATH  Google Scholar 

  8. Chakalov, L.: Sur un probleme de D. Pompeiu. Annuaire Univ. Sofia Phys. Math. Livre I 40, 1–44 (1944)

  9. El Soufi, A., Ilias, S.: Domain deformations and eigenvalues of the Dirichlet Laplacian in Riemannian manifolds. Ill. J. Math. 51, 645–666 (2007)

    MATH  Google Scholar 

  10. Garabedian, P.R., Schiffer, M.: Variational problems in the theory of elliptic partial differential equations. J. Rat. Mech. Anal. 2, 137–171 (1953)

    MathSciNet  MATH  Google Scholar 

  11. Kinderleher, D., Nirenberg, L.: Regularity in free boundary problems. Ann. Scuola Norm. Sup. Pisa Cl. Sci. 4, 373–391 (1977)

    MathSciNet  MATH  Google Scholar 

  12. Kumaresan, S., Prajapat, J.: Serrin’s result for hyperbolic space and sphere. Duke Math. J. 91(1), 17–28 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  13. Liu, G.: Symmetry theorems for the overdetermined eigenvalue problems. J. Differ. Equ. 233, 585–600 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  14. Magnanini, R., Sakaguchi, S.: Heat conductors with a stationary isothermic surface. Ann. Math. Second Ser. 156(3), 931–946 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  15. Molzon, R.: Symmetry and overdetermined boundary problems. Forum Math. 3, 143–156 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  16. Münzner, H.F.: Isoparametrische hyperflachen in spharen I, II. Math. Ann. 251, 57–71 (1980) and 256 (1981), 215–232

  17. Münzner, H.F.: Isoparametrische hyperflachen in spharen I, II. Math. Ann. 256, 215–232 (1981)

  18. Pacard, F., Sicbaldi, P.: Extremal domains for the first eigenvalue of the Laplace–Beltrami operator. Ann. Institut Fourier 59(2), 515–542 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  19. Raulot, S., Savo, A.: On the first eigenvalue of the Dirichlet-to-Neumann operator on forms. J. Funct. Anal. 262, 889–914 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  20. Ros, A., Sicbaldi, P.: Geometry and topology of some overdetermined elliptic problems. J. Differ. Equ. 255, 951–977 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  21. Savo, A.: Uniform estimates and the whole asymptotic series of the heat content on manifolds. Geom. Dedicata 73, 181–214 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  22. Savo, A.: Asymptotics of the heat flow for domains with smooth boundary. Commun. Anal. Geom. 12(3), 671–702 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  23. Schiffer, M.: Hadamard’s formula and variation of domain functions. Am. J. Math. 68, 417–448 (1946)

    Article  MathSciNet  MATH  Google Scholar 

  24. Serrin, J.: A symmetry problem in potential theory. Arch. Rat. Mech. Anal. 43, 304–318 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  25. Shklover, V.: Schiffer problem and isoparametric hypersurfaces. Revista Mat. Iberoamericana 16(3), 529–569 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  26. Tamm, U.: Some aspects of Hankel matrices in coding theory and combinatorics. Electron. J. Combin. 8, # A1 (2001)

  27. Tang, Z., Yan, W.: Isoparametric foliation and Yau conjecture on the first eigenvalue. J. Differ. Geom. 94, 521–540 (2013)

    MathSciNet  MATH  Google Scholar 

  28. Thorbergsson, G.: A survey on isoparametric hypersurfaces and their generalizations. In: Handbook of Differential Geometry, vol. 1. North-Holland, Amsterdam, pp 963–995 (2000)

  29. Wang, Q.M.: Isoparametric functions on Riemannian manifolds, I. Math. Ann. 277, 639–646 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  30. Weinberger, H.F.: Remark on the preceeding paper by Serrin. Arch. Rat. Mech. Anal. 43, 319–320 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  31. Williams, S.A.: A partial solution of the Pompeiu problem. Math. Ann. 223, 183–190 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  32. Williams, S.A.: Analyticity of the boundary for Lipschitz domains without the Pompeiu property. Indiana Univ. Math. J. 30, 357–369 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  33. van den Berg, M., Gilkey, P.: Heat content asymptotics for a Riemannian manifold with boundary. J. Funct. Anal. 120, 48–71 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  34. van den Berg, M., Gilkey, P.: The heat equation with inhomogeneous Dirichlet boundary conditions. Commun. Anal. Geom. 7(2), 279–294 (1999)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

We wish to thank Sylvestre Gallot for his interest in this paper and for many enlightening discussions, and Stefano Capparelli for pointing out the role of orthogonal polynomials in the combinatorial part of the paper (Lemma 27). Finally, we are deeply grateful to the referee, who has done a great job in suggesting a number of remarks which really helped to improve the clarity of the exposition.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alessandro Savo.

Appendix: Proof of Lemma 27

Appendix: Proof of Lemma 27

This section is based on the paper by Ulrich Tamm [26], and we will follow closely the notation there.

Given a sequence \(\{c_0,c_1,c_2,\dots \}\) form the infinite matrix

$$\begin{aligned} A=\begin{pmatrix}c_0&{}c_1&{}c_2&{}\cdots \\ c_1&{}c_2&{}c_3&{}\cdots \\ c_2&{}c_3&{}c_4&{}\cdots \\ \vdots &{}\vdots &{}\vdots &{}\ddots \\ \end{pmatrix}\end{aligned}$$
(68)

and let \(A_n\) be the \(n\times n\) sub-matrix in the upper left corner:

$$\begin{aligned} A_1=(c_0), \quad A_2=\begin{pmatrix} c_0&{}c_1\\ c_1&{}c_2\end{pmatrix}, \quad A_3=\begin{pmatrix} c_0&{}c_1&{}c_2\\ c_1&{}c_2&{}c_3\\ c_2&{}c_3&{}c_4\end{pmatrix}, \dots \end{aligned}$$

In this way we obtain a sequence of Hankel matrices. We will also consider shifted sequences, namely for all \(k\ge 0\) we set

$$\begin{aligned} A^{(k)}=\begin{pmatrix}c_k&{}c_{k+1}&{}c_{k+2}&{}\cdots \\ c_{k+1}&{}c_{k+2}&{}c_{k+3}&{}\cdots \\ c_{k+2}&{}c_{k+3}&{}c_{k+4}&{}\cdots \\ \vdots &{}\vdots &{}\vdots &{}\ddots \\ \end{pmatrix}\end{aligned}$$

and let \(A^{(k)}_n\) be the \(n\times n\) sub-matrix in the upper left corner of \(A^{(k)}\); by convention \(A^{(0)}=A\). For all \(k=0,1,\dots \) we set

$$\begin{aligned} d_n^{(k)}=\det A^{(k)}_n. \end{aligned}$$

By definition, the sequence \(\{d_1^{(k)}, d_2^{(k)}, d_3^{(k)}, \dots \}\) is called the Hankel transform of the sequence \(\{c_k,c_{k+1}, c_{k+2}\dots \}\).

Our interest is in the sequence

$$\begin{aligned} c_m=\left( {\begin{array}{c}2m+1\\ m\end{array}}\right) , \quad m\ge 0. \end{aligned}$$
(69)

It is just the sequence defined in (54) with offset at \(m=0\):

$$\begin{aligned} c_m=a_{m+1}=\{1,3,10,35,126,462,1716, 6735, \dots \}. \end{aligned}$$

Then

$$\begin{aligned} A^{(0)}=\begin{pmatrix}1&{}3&{}10&{}35&{}\cdots \\ 3&{}10&{}35&{}126&{}\cdots \\ 10&{}35&{}126&{}462&{}\cdots \\ \vdots &{}\vdots &{}\vdots &{}\vdots &{}\ddots \\ \end{pmatrix}, \quad \text {and}\quad A^{(1)}=\begin{pmatrix}3&{}10&{}35&{}126&{}\cdots \\ 10&{}35&{}126&{}462&{}\cdots \\ 35&{}126&{}462&{}1716&{}\cdots \\ \vdots &{}\vdots &{}\vdots &{}\vdots &{}\ddots \\ \end{pmatrix}. \end{aligned}$$

From part b) of Proposition 2.1 in [26] we see that, for all n:

$$\begin{aligned} d_n^{(0)}=1, \quad d_n^{(1)}=2n+1, \end{aligned}$$
(70)

and the following expression holds for \(k\ge 2\):

$$\begin{aligned} d_n^{(k)}=\Pi _{1\le i\le j\le k}\dfrac{i+j-1+2n}{i+j-1}. \end{aligned}$$

Now introduce an indeterminate x and consider the infinite matrices

$$\begin{aligned} \mathcal{A}(x)=\begin{pmatrix}1&{}x&{}x^2&{}\cdots \\ c_0&{}c_1&{}c_2&{}\cdots \\ c_1&{}c_2&{}c_3&{}\cdots \\ c_2&{}c_3&{}c_4&{}\cdots \\ \vdots &{}\vdots &{}\vdots &{}\ddots \\ \end{pmatrix}, \quad \mathcal{A}^{(1)}(x)=\begin{pmatrix}1&{}x&{}x^2&{}\cdots \\ c_1&{}c_2&{}c_3&{}\cdots \\ c_2&{}c_3&{}c_4&{}\cdots \\ c_3&{}c_4&{}c_5&{}\cdots \\ \vdots &{}\vdots &{}\vdots &{}\ddots \\ \end{pmatrix}\end{aligned}$$

and the sequences of polynomials \(\{P_n(x)\}, \{P_n^{(1)}(x)\}\) defined by

$$\begin{aligned} \left\{ \begin{aligned}&P_0(x)=1, \quad P_1(x)=\begin{vmatrix} 1&x\\c_0&c_1\end{vmatrix}, \quad P_2(x)=\begin{vmatrix} 1&x&x^2\\c_0&c_1&c_2\\c_1&c_2&c_3\end{vmatrix}, \quad ...\\&P_0^{(1)}(x)=1, \quad P_1^{(1)}(x)=\begin{vmatrix} 1&x\\c_1&c_2\end{vmatrix}, \quad P_2^{(1)}(x)=\begin{vmatrix} 1&x&x^2\\c_1&c_2&c_3\\c_2&c_3&c_4\end{vmatrix}, \quad ...\end{aligned}\right. \end{aligned}$$

Recalling that \(c_m=a_{m+1}\) we see that proving Lemma 27 amounts to prove that, for all n:

$$\begin{aligned} P_n(4)=(-1)^n, \quad P_n^{(1)}(4)=(-1)^n(n+1). \end{aligned}$$
(71)

Note that, by (70), the leading coefficient of \(P_n(x)\) (resp. \(P_n^{(1)}(x)\)) is \((-1)^{n}d_n=(-1)^n\) (resp. \((-1)^nd_n^{(1)}=(-1)^n(2n+1)\)). Then, the polynomials

$$\begin{aligned} t_n(x)=(-1)^{n}P_n(x), \quad t_n^{(1)}(x)=\dfrac{(-1)^n}{2n+1}P_n^{(1)}(x) \end{aligned}$$
(72)

are of degree n and monic. Taking into account (71) and (72), to prove Lemma 27 it is then sufficient to show the following fact.

Lemma 29

For the polynomials defined in (72) one has, for all n:

$$\begin{aligned} t_n(4)=1,\quad t_n^{(1)}(4)=\dfrac{n+1}{2n+1}. \end{aligned}$$

To prove Lemma 29, we use the theory of orthogonal polynomials. We first observe:

Lemma 30

The polynomials \(\{t_n(x)\}\) defined in (72) coincide with the polynomials defined in equations (1.7) and (1.8) in [26]. They are orthogonal with respect to the linear operator T mapping \(x^k\) to its moment \(c_k\) for all k: i.e.

$$\begin{aligned} T(t_m(x)\cdot t_n(x))=0\quad \text {for all } m\ne n. \end{aligned}$$

Similarly, the polynomials \(\{t_n^{(1)}(x)\}\) are orthogonal with respect to the shifted linear operator mapping \(x^k\) to \(c_{k+1}\) for all k.

Proof

We closely follow the exposition in [26], page 3, with a slight change of notation to suit our needs. For all \(n\ge 1\), consider the following equation in the unknowns \(\alpha _{n,0},\dots ,\alpha _{n,n-1}\):

$$\begin{aligned} \begin{pmatrix}c_0&{}c_1&{}c_2&{}\cdots &{}c_{n-1}\\ c_1&{}c_2&{}c_3&{}\cdots &{}c_{n}\\ c_2&{}c_3&{}c_4&{}\cdots &{}c_{n+1}\\ \ \vdots &{}\vdots &{}\vdots &{}\ddots &{}\vdots \\ c_{n-1}&{}c_n&{}c_{n+1}&{}\cdots &{}c_{2n-2}\\ \end{pmatrix}\cdot \begin{pmatrix} \alpha _{n,0}\\ \alpha _{n,1}\\ \alpha _{n,2}\\ \vdots \\ \alpha _{n,n-1}\end{pmatrix}=\begin{pmatrix} -c_n\\ -c_{n+1}\\ -c_{n+2}\\ \vdots \\ -c_{2n-1}\end{pmatrix}. \end{aligned}$$
(73)

In our case, we know that \(\det A_n=1\) for all n, hence \(A_n\) is non-singular and (73) has a unique solution. As in (1.8) of [26], define, for \(n\ge 1\), the monic polynomials

$$\begin{aligned} \tau _n(x)\doteq x^n+\alpha _{n,n-1}x^{n-1}+\alpha _{n,n-2}x^{n-2}+\dots +\alpha _{n,1}x+\alpha _{n,0}. \end{aligned}$$
(74)

Now, by Cramer’s rule, \(\alpha _{n,k}\) is the determinant of the matrix obtained by replacing the \((k+1)\)-st column of \(A_n\) by the column in the right-hand side of (73); in turn, this determinant is also equal to \((-1)^n\) times the coefficient of \(x^k\) in \(P_n(x)\). This means that

$$\begin{aligned} \tau _n(x)=t_n(x) \end{aligned}$$

as asserted. The orthogonality of the sequence \(\{t_n(x)\}\) with respect to the linear mapping T has been proved in a book by Bressoud (as cited in [26]); however it is not difficult to verify that, from definition (72) of \(t_n(x)\), one has \(T(x^m\cdot t_n(x))=0\) for all \(m=0,\dots ,n-1\), which easily implies this fact. A similar argument applies to the polynomials \(t_n^{(1)}(x)\) associated to the shifted sequence. \(\square \)

Proof of Lemma 29 Orthogonal polynomials satisfy a three-term recursive relation: following [26], pages 5–6 and equation (1.19), one can write:

$$\begin{aligned} \left\{ \begin{aligned}&t_n(x)=(x-\alpha _n)t_{n-1}(x)-\beta _{n-1}t_{n-2}(x), \quad t_0(x)=1, t_1(x)=x-\alpha _1\\&t_n^{(1)}(x)=(x-\alpha _n^{(1)})t_{n-1}^{(1)}(x)-\beta _{n-1}^{(1)}t_{n-2}^{(1)}(x) \quad t_0^{(1)}(x)=1, t_1^{(1)}(x)=x-\alpha _1^{(1)}\end{aligned}\right. \end{aligned}$$
(75)

for suitable numerical sequences \(\alpha _n\doteq \alpha _n^{(0)},\beta _n\doteq \beta _n^{(0)}, \alpha _n^{(1)}, \beta _{n-1}^{(1)}\). In fact, for arbitrary k, the coefficients \(\alpha _n^{(k)}, \beta _{n-1}^{(k)}\) can be computed in terms of the coefficients \(q_n^{(k)},e_n^{(k)}\) in the continued fraction expansion of \(1-xF(x)\), where \(F(x)=\sum _{m=0}^{\infty }c_{m+k}x^m\) (see (1.14), (1.19) and (1.20) and (1.21) in [26]):

$$\begin{aligned} \alpha _1=q_1\quad \text {and, for } n\ge 1: \quad \alpha _{n+1}^{(k)}=q_{n+1}^{(k)}+e_{n}^{(k)},\quad \beta _n^{(k)}=q_n^{(k)}\cdot e_n^{(k)}, \end{aligned}$$
(76)

For our sequence \(c_m=\left( {\begin{array}{c}2m+1\\ m\end{array}}\right) \) the corresponding coefficients have been computed in the equation (2.6) of Corollary 2.1 of [26].

$$\begin{aligned} q_n^{(k)}=\dfrac{(2n+2k)(2n+2k+1)}{(2n+k-1)(2n+k)},\quad e_n^{(k)}=\dfrac{(2n-1)(2n)}{(2n+k)(2n+k+1)}. \end{aligned}$$

In particular,

$$\begin{aligned} \left\{ \begin{aligned}&q_n=q_n^{(0)}=\dfrac{2n+1}{2n-1}\\&e_n=e_n^{(0)}=\dfrac{2n-1}{2n+1}\end{aligned}\right. \quad \text {and}\quad \left\{ \begin{aligned}&q_n^{(1)}=\dfrac{(2n+2)(2n+3)}{(2n)(2n+1)}\\&e_n^{(1)}=\dfrac{(2n-1)(2n)}{(2n+1)(2n+2)}.\end{aligned}\right. \end{aligned}$$

From (76) we obtain \(\alpha _1=q_1=3\), \(\alpha _1^{(1)}= q_1^{(1)}=\frac{10}{3}\) and for \(n\ge 2\) (after some calculations):

$$\begin{aligned} \left\{ \begin{aligned}&\alpha _n=2\\&\beta _{n-1}=1\end{aligned}\right. , \quad \text {and}\quad \left\{ \begin{aligned}&\alpha _n^{(1)}=2+\frac{4}{(2n+1)(2n-1)}\\&\beta _{n-1}^{(1)}=\dfrac{(2n-3)(2n+1)}{(2n-1)^2}\end{aligned}\right. \end{aligned}$$
(77)

From (75) and (77) one has the following recursive scheme for the sequence \(\{t_n(x)\}\):

$$\begin{aligned} \left\{ \begin{aligned}&t_n(x)=(x-2)t_{n-1}(x)-t_{n-2}(x)\\&t_0(x)=1\\ {}&t_1(x)=x-3\end{aligned}\right. \end{aligned}$$

Setting \(X_n=t_n(4)\) we obtain

$$\begin{aligned} \left\{ \begin{aligned}&X_n=2X_{n-1}-X_{n-2}\\&X_0=1\\ {}&X_1=1\end{aligned}\right. \end{aligned}$$

hence \(X_n=1\) for all n. That is:

$$\begin{aligned} t_n(4)=1 \end{aligned}$$

for all n. This shows the first relation in the Lemma.

About the sequence \(\{t_n^{(1)}(x)\}\) we know:

$$\begin{aligned} \left\{ \begin{aligned}&t_n^{(1)}(x)=(x-\alpha _n^{(1)})t_{n-1}^{(1)}(x)-\beta _{n-1}^{(1)}t_{n-2}^{(1)}(x)\\&t_0^{(1)}(x)=1\\ {}&t_1^{(1)}(x)=x-\frac{10}{3}\end{aligned}\right. \end{aligned}$$

where \(\alpha _n^{(1)}\) and \(\beta _{n-1}^{(1)}\) are as in (77). We are interested in the sequence \( X_n=t_n^{(1)}(4), \) which obeys the recursive law:

$$\begin{aligned} \left\{ \begin{aligned}&X_n=(4-\alpha _n^{(1)})X_{n-1}-\beta _{n-1}^{(1)}X_{n-2}\\&X_0=1\\ {}&X_1=\frac{2}{3}\end{aligned}\right. \end{aligned}$$

By induction on n, let us show that \(X_n=\frac{n+1}{2n+1}\). In fact this holds for \(n=1\). Assume it to be true for all \(k\le n-1\). By (77):

$$\begin{aligned} 4-\alpha _n^{(1)}=\dfrac{8n^2-6}{(2n-1)(2n+1)}, \quad \beta _{n-1}^{(1)}=\dfrac{(2n-3)(2n+1)}{(2n-1)^2} \end{aligned}$$

hence:

$$\begin{aligned} X_n&=(4-\alpha _n^{(1)})X_{n-1}-\beta _{n-1}^{(1)}X_{n-2}\\&=\dfrac{8n^2-6}{(2n-1)(2n+1)}\cdot \frac{n}{2n-1}-\dfrac{(2n-3)(2n+1)}{(2n-1)^2}\cdot \frac{n-1}{2n-3}\\&=\frac{n+1}{2n+1} \end{aligned}$$

which shows the claim. With this, the proof of Lemma 27 is complete.

Remark We point out that the expression of \(\alpha _{n+1}^{(k)}\) in the statement of Corollary 2.3 in [26] is wrong; this is due to an incorrect algebraic manipulation of \(q_{n+1}^{(k)}+e_n^{(k)}\) in the proof of Corollary 2.3 there. The correct value of \(\alpha _{n+1}^{(1)}\) shown in (77) is directly computed from the expressions of \(q_n^{(1)}\) and \(e_n^{(1)}\) taken from Corollary 2.1, equation (2.6) of [26].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Savo, A. Heat flow, heat content and the isoparametric property. Math. Ann. 366, 1089–1136 (2016). https://doi.org/10.1007/s00208-015-1359-9

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00208-015-1359-9

Mathematics Subject Classification

Navigation