Riesz–Thorin theorem
   HOME

TheInfoList



OR:

In
mathematics Mathematics is an area of knowledge that includes the topics of numbers, formulas and related structures, shapes and the spaces in which they are contained, and quantities and their changes. These topics are represented in modern mathematics ...
, the Riesz–Thorin theorem, often referred to as the Riesz–Thorin interpolation theorem or the Riesz–Thorin convexity theorem, is a result about ''interpolation of operators''. It is named after Marcel Riesz and his student
G. Olof Thorin G is the seventh letter of the Latin alphabet. G may also refer to: Places * Gabon, international license plate code G * Glasgow, UK postal code G * Eastern Quebec, Canadian postal prefix G * Melbourne Cricket Ground in Melbourne, Australia, ...
. This theorem bounds the norms of linear maps acting between spaces. Its usefulness stems from the fact that some of these spaces have rather simpler structure than others. Usually that refers to which is a
Hilbert space In mathematics, Hilbert spaces (named after David Hilbert) allow generalizing the methods of linear algebra and calculus from (finite-dimensional) Euclidean vector spaces to spaces that may be infinite-dimensional. Hilbert spaces arise natural ...
, or to and . Therefore one may prove theorems about the more complicated cases by proving them in two simple cases and then using the Riesz–Thorin theorem to pass from the simple cases to the complicated cases. The
Marcinkiewicz theorem In mathematics, the Marcinkiewicz interpolation theorem, discovered by , is a result bounding the norms of non-linear operators acting on ''L''p spaces. Marcinkiewicz' theorem is similar to the Riesz–Thorin theorem about linear operators, bu ...
is similar but applies also to a class of non-linear maps.


Motivation

First we need the following definition: :Definition. Let be two numbers such that . Then for define by: . By splitting up the function in as the product and applying Hölder's inequality to its power, we obtain the following result, foundational in the study of -spaces: This result, whose name derives from the convexity of the map on , implies that . On the other hand, if we take the ''layer-cake decomposition'' , then we see that and , whence we obtain the following result: In particular, the above result implies that is included in , the
sumset In additive combinatorics, the sumset (also called the Minkowski sum) of two subsets A and B of an abelian group G (written additively) is defined to be the set of all sums of an element from A with an element from B. That is, :A + B = \. The n-f ...
of and in the space of all measurable functions. Therefore, we have the following chain of inclusions: In practice, we often encounter
operators Operator may refer to: Mathematics * A symbol indicating a mathematical operation * Logical operator or logical connective in mathematical logic * Operator (mathematics), mapping that acts on elements of a space to produce elements of another sp ...
defined on the
sumset In additive combinatorics, the sumset (also called the Minkowski sum) of two subsets A and B of an abelian group G (written additively) is defined to be the set of all sums of an element from A with an element from B. That is, :A + B = \. The n-f ...
. For example, the
Riemann–Lebesgue lemma In mathematics, the Riemann–Lebesgue lemma, named after Bernhard Riemann and Henri Lebesgue, states that the Fourier transform or Laplace transform of an ''L''1 function vanishes at infinity. It is of importance in harmonic analysis and asymptot ...
shows that the
Fourier transform A Fourier transform (FT) is a mathematical transform that decomposes functions into frequency components, which are represented by the output of the transform as a function of frequency. Most commonly functions of time or space are transformed ...
maps boundedly into , and Plancherel's theorem shows that the Fourier transform maps boundedly into itself, hence the Fourier transform \mathcal extends to by setting \mathcal(f_1+f_2) = \mathcal_(f_1) + \mathcal_(f_2) for all and . It is therefore natural to investigate the behavior of such operators on the ''intermediate subspaces'' . To this end, we go back to our example and note that the Fourier transform on the sumset was obtained by taking the sum of two instantiations of the same operator, namely \mathcal_:L^1(\mathbf^d) \to L^\infty(\mathbf^d), \mathcal_:L^2(\mathbf^d) \to L^2(\mathbf^d). These really are the ''same'' operator, in the sense that they agree on the subspace . Since the intersection contains simple functions, it is dense in both and . Densely defined continuous operators admit unique extensions, and so we are justified in considering \mathcal_ and \mathcal_ to be ''the same''. Therefore, the problem of studying operators on the sumset essentially reduces to the study of operators that map two natural domain spaces, and , boundedly to two target spaces: and , respectively. Since such operators map the sumset space to , it is natural to expect that these operators map the intermediate space to the corresponding intermediate space .


Statement of the theorem

There are several ways to state the Riesz–Thorin interpolation theorem; to be consistent with the notations in the previous section, we shall use the sumset formulation. In other words, if is simultaneously of type and of type , then is of type for all . In this manner, the interpolation theorem lends itself to a pictorial description. Indeed, the Riesz diagram of is the collection of all points in the unit square such that is of type . The interpolation theorem states that the Riesz diagram of is a convex set: given two points in the Riesz diagram, the line segment that connects them will also be in the diagram. The interpolation theorem was originally stated and proved by Marcel Riesz in 1927. The 1927 paper establishes the theorem only for the ''lower triangle'' of the Riesz diagram, viz., with the restriction that and .
Olof Thorin Olov (or Olof) is a Swedish form of Olav/Olaf, meaning "ancestor's descendant". A common short form of the name is ''Olle''. The name may refer to: *Per-Olov Ahrén (1926–2004), Swedish clergyman, bishop of Lund from 1980 to 1992 * Per-Olov Br ...
extended the interpolation theorem to the entire square, removing the lower-triangle restriction. The proof of Thorin was originally published in 1938 and was subsequently expanded upon in his 1948 thesis.


Proof

We will first prove the result for simple functions and eventually show how the argument can be extended by density to all measurable functions.


Simple Functions

By symmetry, let us assume p_0 < p_1 (the case p_0 = p_1 trivially follows from ()). Let f be a simple function, that is f = \sum_^m a_j \mathbf_ for some finite m\in\mathbb, a_j = \left\vert a_j\right\vert\mathrm^ \in \mathbb and A_j\in\Sigma_1, j=1,2,\dots,m. Similarly, let g denote a simple function \Omega_2 \to \mathbb, namely g = \sum_^n b_k \mathbf_ for some finite n\in\mathbb, b_k = \left\vert b_k\right\vert\mathrm^ \in \mathbb and B_k\in\Sigma_2, k=1,2,\dots,n. Note that, since we are assuming \Omega_1 and \Omega_2 to be \sigma-finite metric spaces, f\in L^(\mu_1) and g\in L^r(\mu_2) for all r \in , \infty/math>. Then, by proper normalization, we can assume \lVert f\rVert_= 1 and \lVert g\rVert_=1, with q_\theta' = q_\theta(q_\theta-1)^ and with p_\theta, q_\theta as defined by the theorem statement. Next, we define the two complex functions \begin u: \mathbb&\to \mathbb& v: \mathbb&\to \mathbb\\ z &\mapsto u(z)=\frac + \frac & z &\mapsto v(z)=\frac + \frac.\end Note that, for z=\theta, u(\theta) = p_\theta^ and v(\theta) = q_\theta^. We then extend f and g to depend on a complex parameter z as follows: \begin f_z &= \sum_^m \left\vert a_j\right\vert^ \mathrm^ \mathbf_ \\ g_z &= \sum_^n \left\vert b_k\right\vert^ \mathrm^ \mathbf_\end so that f_\theta = f and g_\theta = g. Here, we are implicitly excluding the case q_0 = q_1 = 1, which yields v\equiv 1: In that case, one can simply take g_z=g, independently of z, and the following argument will only require minor adaptations. Let us now introduce the function \Phi(z) = \int_ (T f_z) g_z \,\mathrm\mu_2 = \sum_^m \sum_^n \left\vert a_j\right\vert^ \left\vert b_k\right\vert^ \gamma_ where \gamma_ = \mathrm^ \int_ (T \mathbf_) \mathbf_ \,\mathrm\mu_2 are constants independent of z. We readily see that \Phi(z) is an entire function, bounded on the strip 0 \le \operatornamez \le 1. Then, in order to prove (), we only need to show that for all f_z and g_z as constructed above. Indeed, if () holds true, by
Hadamard three-lines theorem In complex analysis, a branch of mathematics, the Hadamard three-lines theorem is a result about the behaviour of holomorphic functions defined in regions bounded by parallel lines in the complex plane. The theorem is named after the French mathem ...
, \left\vert\Phi(\theta + \mathrm0)\right\vert = \biggl\vert\int_ (Tf) g \,\mathrm\mu_2\biggr\vert \le \, T\, _^ \, T\, _^\theta for all f and g. This means, by fixing f, that \sup_g \biggl\vert\int_ (Tf) g \,\mathrm\mu_2\biggr\vert \le \, T\, _^ \, T\, _^\theta where the supremum is taken with respect to all g simple functions with \lVert g\rVert_ = 1. The left-hand side can be rewritten by means of the following lemma. In our case, the lemma above implies \lVert Tf\rVert_ \le \, T\, _^ \, T\, _^\theta for all simple function f with \lVert f\rVert_ = 1. Equivalently, for a generic simple function, \lVert Tf\rVert_ \le \, T\, _^ \, T\, _^\theta \lVert f\rVert_.


Proof of ()

Let us now prove that our claim () is indeed certain. The sequence (A_j)_^m consists of disjoint subsets in \Sigma_1 and, thus, each \xi\in \Omega_1 belongs to (at most) one of them, say A_. Then, for z=\mathrmy, \begin \left\vert f_(\xi)\right\vert &= \left\vert a_\right\vert^\frac \\ &= \exp\biggl(\log\left\vert a_\right\vert\frac\biggr) \exp\biggl(-\mathrmy \log\left\vert a_\right\vert p_\theta\biggl(\frac - \frac \biggr) \biggr) \\ &= \left\vert a_\right\vert^ \\ & = \left\vert f(\xi)\right\vert^\end which implies that \lVert f_\rVert_ \le \lVert f\rVert_^. With a parallel argument, each \zeta \in \Omega_2 belongs to (at most) one of the sets supporting g, say B_, and \left\vert g_(\zeta)\right\vert = \left\vert b_\right\vert^ = \left\vert g(\zeta)\right\vert^ = \left\vert g(\zeta)\right\vert^ \implies \lVert g_\rVert_ \le \lVert g\rVert_^. We can now bound \Phi(\mathrmy): By applying Hölder’s inequality with conjugate exponents q_0 and q_0', we have \begin \left\vert\Phi(\mathrmy)\right\vert &\le \lVert T f_\rVert_ \lVert g_\rVert_ \\ &\le \, T\, _ \lVert f_\rVert_ \lVert g_\rVert_ \\ &= \, T\, _ \lVert f\rVert_^ \lVert g\rVert_^ \\ &= \, T\, _.\end We can repeat the same process for z=1+\mathrmy to obtain \left\vert f_(\xi)\right\vert = \left\vert f(\xi)\right\vert^, \left\vert g_(\zeta)\right\vert = \left\vert g(\zeta)\right\vert^ and, finally, \left\vert\Phi(1+\mathrmy)\right\vert \le \, T\, _ \lVert f_\rVert_ \lVert g_\rVert_ = \, T\, _.


Extension to All Measurable Functions in L^

So far, we have proven that when f is a simple function. As already mentioned, the inequality holds true for all f\in L^(\Omega_1) by the density of simple functions in L^(\Omega_1). Formally, let f\in L^(\Omega_1) and let (f_n)_n be a sequence of simple functions such that \left\vert f_n\right\vert \le \left\vert f\right\vert, for all n, and f_n \to f pointwise. Let E=\ and define g = f \mathbf_E, g_n = f_n \mathbf_E, h = f - g = f \mathbf_ and h_n = f_n - g_n. Note that, since we are assuming p_0 \le p_\theta \le p_1, \begin \lVert f\rVert_^ &= \int_ \left\vert f\right\vert^ \,\mathrm\mu_1 \ge \int_ \left\vert f\right\vert^ \mathbf_ \,\mathrm\mu_1 \ge \int_ \left\vert f \mathbf_\right\vert^ \,\mathrm\mu_1 = \int_ \left\vert g\right\vert^ \,\mathrm\mu_1 = \lVert g\rVert_^ \\ \lVert f\rVert_^ &= \int_ \left\vert f\right\vert^ \,\mathrm\mu_1 \ge \int_ \left\vert f\right\vert^ \mathbf_ \,\mathrm\mu_1 \ge \int_ \left\vert f \mathbf_\right\vert^ \,\mathrm\mu_1 = \int_ \left\vert h\right\vert^ \,\mathrm\mu_1 = \lVert h\rVert_^\end and, equivalently, g\in L^(\Omega_1) and h\in L^(\Omega_1). Let us see what happens in the limit for n\to\infty. Since \left\vert f_n\right\vert \le \left\vert f\right\vert, \left\vert g_n\right\vert \le \left\vert g\right\vert and \left\vert h_n\right\vert \le \left\vert h\right\vert, by the dominated convergence theorem one readily has \begin \lVert f_n\rVert_ &\to \lVert f\rVert_ & \lVert g_n\rVert_ &\to \lVert g\rVert_ & \lVert h_n\rVert_ &\to \lVert h\rVert_.\end Similarly, \left\vert f - f_n\right\vert \le 2\left\vert f\right\vert, \left\vert g-g_n\right\vert \le 2\left\vert g\right\vert and \left\vert h - h_n\right\vert \le 2\left\vert h\right\vert imply \begin \lVert f - f_n\rVert_ &\to 0 & \lVert g - g_n\rVert_ &\to 0 & \lVert h - h_n\rVert_ &\to 0\end and, by the linearity of T as an operator of types (p_0, q_0) and (p_1, q_1) (we have not proven yet that it is of type (p_\theta, q_\theta) for a generic f) \begin \lVert Tg - Tg_n\rVert_ & \le \, T\, _ \lVert g - g_n\rVert_ \to 0 & \lVert Th - Th_n\rVert_ & \le \, T\, _ \lVert h - h_n\rVert_ \to 0.\end It is now easy to prove that Tg_n \to Tg and Th_n \to Th in measure: For any \epsilon > 0,
Chebyshev’s inequality In probability theory, Chebyshev's inequality (also called the Bienaymé–Chebyshev inequality) guarantees that, for a wide class of probability distributions, no more than a certain fraction of values can be more than a certain distance from th ...
yields \mu_2(y\in \Omega_2: \left\vert Tg - Tg_n\right\vert > \epsilon) \le \frac and similarly for Th - Th_n. Then, Tg_n \to Tg and Th_n \to Th a.e. for some subsequence and, in turn, Tf_n \to Tf a.e. Then, by
Fatou’s lemma In mathematics, Fatou's lemma establishes an inequality relating the Lebesgue integral of the limit inferior of a sequence of functions to the limit inferior of integrals of these functions. The lemma is named after Pierre Fatou. Fatou's lem ...
and recalling that () holds true for simple functions, \lVert Tf\rVert_ \le \liminf_ \lVert T f_n\rVert_ \le \, T\, _ \liminf_ \lVert f_n\rVert_ = \, T\, _ \lVert f\rVert_.


Interpolation of analytic families of operators

The proof outline presented in the above section readily generalizes to the case in which the operator is allowed to vary analytically. In fact, an analogous proof can be carried out to establish a bound on the entire function \varphi(z) = \int (T_z f_z)g_z \, d\mu_2, from which we obtain the following theorem of Elias Stein, published in his 1956 thesis: The theory of real Hardy spaces and the space of bounded mean oscillations permits us to wield the Stein interpolation theorem argument in dealing with operators on the Hardy space and the space of bounded mean oscillations; this is a result of
Charles Fefferman Charles Louis Fefferman (born April 18, 1949) is an American mathematician at Princeton University, where he is currently the Herbert E. Jones, Jr. '43 University Professor of Mathematics. He was awarded the Fields Medal in 1978 for his contri ...
and Elias Stein.


Applications


Hausdorff–Young inequality

It has been shown in the first section that the
Fourier transform A Fourier transform (FT) is a mathematical transform that decomposes functions into frequency components, which are represented by the output of the transform as a function of frequency. Most commonly functions of time or space are transformed ...
\mathcal maps boundedly into and into itself. A similar argument shows that the Fourier series operator, which transforms periodic functions into functions \hat:\mathbf \to \mathbf whose values are the Fourier coefficients \hat(n) = \frac \int_^ f(x) e^ \, dx , maps boundedly into and into . The Riesz–Thorin interpolation theorem now implies the following: \begin \left \, \mathcalf \right \, _ &\leq \, f\, _ \\ \left \, \hat \right \, _ &\leq \, f\, _ \end where and . This is the
Hausdorff–Young inequality The Hausdorff−Young inequality is a foundational result in the mathematical field of Fourier analysis. As a statement about Fourier series, it was discovered by and extended by . It is now typically understood as a rather direct corollary of th ...
. The Hausdorff–Young inequality can also be established for the Fourier transform on locally compact Abelian groups. The norm estimate of 1 is not optimal. See the main article for references.


Convolution operators

Let be a fixed integrable function and let be the operator of convolution with , i.e., for each function we have . It is well known that is bounded from to and it is trivial that it is bounded from to (both bounds are by ). Therefore the Riesz–Thorin theorem gives \, f * g \, _p \leq \, f\, _1 \, g\, _p. We take this inequality and switch the role of the operator and the operand, or in other words, we think of as the operator of convolution with , and get that is bounded from to ''Lp''. Further, since is in we get, in view of Hölder's inequality, that is bounded from to , where again . So interpolating we get \, f*g\, _s\leq \, f\, _r\, g\, _p where the connection between ''p'', ''r'' and ''s'' is \frac+\frac=1+\frac.


The Hilbert transform

The Hilbert transform of is given by \mathcalf(x) = \frac \, \mathrm \int_^\infty \frac \, dt = \left(\frac \, \mathrm \frac \ast f\right)(x), where p.v. indicates the
Cauchy principal value In mathematics, the Cauchy principal value, named after Augustin Louis Cauchy, is a method for assigning values to certain improper integrals which would otherwise be undefined. Formulation Depending on the type of singularity in the integrand ...
of the integral. The Hilbert transform is a Fourier multiplier operator with a particularly simple multiplier: \widehat(\xi) = -i \, \sgn(\xi) \hat(\xi). It follows from the
Plancherel theorem In mathematics, the Plancherel theorem (sometimes called the Parseval–Plancherel identity) is a result in harmonic analysis, proven by Michel Plancherel in 1910. It states that the integral of a function's squared modulus is equal to the integ ...
that the Hilbert transform maps boundedly into itself. Nevertheless, the Hilbert transform is not bounded on or , and so we cannot use the Riesz–Thorin interpolation theorem directly. To see why we do not have these endpoint bounds, it suffices to compute the Hilbert transform of the simple functions and . We can show, however, that (\mathcalf)^2 = f^2 + 2\mathcal(f\mathcalf) for all Schwartz functions , and this identity can be used in conjunction with the
Cauchy–Schwarz inequality The Cauchy–Schwarz inequality (also called Cauchy–Bunyakovsky–Schwarz inequality) is considered one of the most important and widely used inequalities in mathematics. The inequality for sums was published by . The corresponding inequality f ...
to show that the Hilbert transform maps boundedly into itself for all . Interpolation now establishes the bound \, \mathcalf\, _p \leq A_p \, f\, _p for all , and the self-adjointness of the Hilbert transform can be used to carry over these bounds to the case.


Comparison with the real interpolation method

While the Riesz–Thorin interpolation theorem and its variants are powerful tools that yield a clean estimate on the interpolated operator norms, they suffer from numerous defects: some minor, some more severe. Note first that the complex-analytic nature of the proof of the Riesz–Thorin interpolation theorem forces the scalar field to be . For extended-real-valued functions, this restriction can be bypassed by redefining the function to be finite everywhere—possible, as every integrable function must be finite almost everywhere. A more serious disadvantage is that, in practice, many operators, such as the Hardy–Littlewood maximal operator and the Calderón–Zygmund operators, do not have good endpoint estimates. Elias Stein is quoted for saying that interesting operators in
harmonic analysis Harmonic analysis is a branch of mathematics concerned with the representation of functions or signals as the superposition of basic waves, and the study of and generalization of the notions of Fourier series and Fourier transforms (i.e. an ex ...
are rarely bounded on and .
In the case of the Hilbert transform in the previous section, we were able to bypass this problem by explicitly computing the norm estimates at several midway points. This is cumbersome and is often not possible in more general scenarios. Since many such operators satisfy the weak-type estimates \mu \left( \ \right) \leq \left( \frac \right)^q, real interpolation theorems such as the
Marcinkiewicz interpolation theorem In mathematics, the Marcinkiewicz interpolation theorem, discovered by , is a result bounding the norms of non-linear operators acting on ''L''p spaces. Marcinkiewicz' theorem is similar to the Riesz–Thorin theorem about linear operators, but ...
are better-suited for them. Furthermore, a good number of important operators, such as the Hardy-Littlewood maximal operator, are only
sublinear In linear algebra, a sublinear function (or functional as is more often used in functional analysis), also called a quasi-seminorm or a Banach functional, on a vector space X is a real-valued function with only some of the properties of a seminorm. ...
. This is not a hindrance to applying real interpolation methods, but complex interpolation methods are ill-equipped to handle non-linear operators. On the other hand, real interpolation methods, compared to complex interpolation methods, tend to produce worse estimates on the intermediate operator norms and do not behave as well off the diagonal in the Riesz diagram. The off-diagonal versions of the Marcinkiewicz interpolation theorem require the formalism of Lorentz spaces and do not necessarily produce norm estimates on the -spaces.


Mityagin's theorem

B. Mityagin extended the Riesz–Thorin theorem; this extension is formulated here in the special case of spaces of sequences with unconditional bases (cf. below). Assume: \, A\, _, \, A\, _ \leq M. Then \, A\, _ \leq M for any unconditional Banach space of sequences , that is, for any (x_i) \in X and any (\varepsilon_i) \in \^\infty, \, (\varepsilon_i x_i) \, _X = \, (x_i) \, _X . The proof is based on the
Krein–Milman theorem In the mathematical theory of functional analysis, the Krein–Milman theorem is a proposition about compact convex sets in locally convex topological vector spaces (TVSs). This theorem generalizes to infinite-dimensional spaces and to arbitrar ...
.


See also

*
Marcinkiewicz interpolation theorem In mathematics, the Marcinkiewicz interpolation theorem, discovered by , is a result bounding the norms of non-linear operators acting on ''L''p spaces. Marcinkiewicz' theorem is similar to the Riesz–Thorin theorem about linear operators, but ...
*
Interpolation space In the field of mathematical analysis, an interpolation space is a space which lies "in between" two other Banach spaces. The main applications are in Sobolev spaces, where spaces of functions that have a noninteger number of derivatives are interpo ...


Notes


References

* . * * . Translated from the Russian and edited by G. P. Barker and G. Kuerti. * . * . * * * * *


External links

* {{DEFAULTSORT:Riesz-Thorin theorem Theorems involving convexity Theorems in harmonic analysis Theorems in Fourier analysis Theorems in functional analysis Banach spaces Operator theory