Hölder's inequality

From HandWiki
Short description: Inequality between integrals in Lp spaces

In mathematical analysis, Hölder's inequality, named after Otto Hölder, is a fundamental inequality between integrals and an indispensable tool for the study of Lp spaces.

Hölder's inequality — Let (S, Σ, μ) be a measure space and let p, q [1, ∞] with 1/p + 1/q = 1. Then for all measurable real- or complex-valued functions f and g on S,

[math]\displaystyle{ \|fg\|_1 \le \|f\|_p \|g\|_q. }[/math]

If, in addition, p, q (1, ∞) and fLp(μ) and gLq(μ), then Hölder's inequality becomes an equality if and only if |f|p and |g|q are linearly dependent in L1(μ), meaning that there exist real numbers α, β ≥ 0, not both of them zero, such that α|f |p = β |g|q μ-almost everywhere.

The numbers p and q above are said to be Hölder conjugates of each other. The special case p = q = 2 gives a form of the Cauchy–Schwarz inequality.[1] Hölder's inequality holds even if ||fg||1 is infinite, the right-hand side also being infinite in that case. Conversely, if f is in Lp(μ) and g is in Lq(μ), then the pointwise product fg is in L1(μ).

Hölder's inequality is used to prove the Minkowski inequality, which is the triangle inequality in the space Lp(μ), and also to establish that Lq(μ) is the dual space of Lp(μ) for p [1, ∞).

Hölder's inequality (in a slightly different form) was first found by Leonard James Rogers (1888). Inspired by Rogers' work, (Hölder 1889) gave another proof as part of a work developing the concept of convex and concave functions and introducing Jensen's inequality,[2] which was in turn named for work of Johan Jensen building on Hölder's work.[3]

Remarks

Conventions

The brief statement of Hölder's inequality uses some conventions.

  • In the definition of Hölder conjugates, 1/∞ means zero.
  • If p, q [1, ∞), then ||f||p and ||g||q stand for the (possibly infinite) expressions
[math]\displaystyle{ \begin{align} &\left(\int_S |f|^p\,\mathrm{d}\mu\right)^{\frac{1}{p}} \\ &\left(\int_S |g|^q\,\mathrm{d}\mu\right)^{\frac{1}{q}} \end{align} }[/math]
  • If p = ∞, then ||f|| stands for the essential supremum of |f|, similarly for ||g||.
  • The notation ||f||p with 1 ≤ p ≤ ∞ is a slight abuse, because in general it is only a norm of f if ||f||p is finite and f is considered as equivalence class of μ-almost everywhere equal functions. If fLp(μ) and gLq(μ), then the notation is adequate.
  • On the right-hand side of Hölder's inequality, 0 × ∞ as well as ∞ × 0 means 0. Multiplying a > 0 with ∞ gives ∞.

Estimates for integrable products

As above, let f and g denote measurable real- or complex-valued functions defined on S. If ||fg||1 is finite, then the pointwise products of f with g and its complex conjugate function are μ-integrable, the estimate

[math]\displaystyle{ \biggl|\int_S f\bar g\,\mathrm{d}\mu\biggr|\le\int_S|fg|\,\mathrm{d}\mu =\|fg\|_1 }[/math]

and the similar one for fg hold, and Hölder's inequality can be applied to the right-hand side. In particular, if f and g are in the Hilbert space L2(μ), then Hölder's inequality for p = q = 2 implies

[math]\displaystyle{ |\langle f,g\rangle| \le \|f\|_2 \|g\|_2, }[/math]

where the angle brackets refer to the inner product of L2(μ). This is also called Cauchy–Schwarz inequality, but requires for its statement that ||f||2 and ||g||2 are finite to make sure that the inner product of f and g is well defined. We may recover the original inequality (for the case p = 2) by using the functions |f| and |g| in place of f and g.

Generalization for probability measures

If (S, Σ, μ) is a probability space, then p, q [1, ∞] just need to satisfy 1/p + 1/q ≤ 1, rather than being Hölder conjugates. A combination of Hölder's inequality and Jensen's inequality implies that

[math]\displaystyle{ \|fg\|_1 \le \|f\|_p \|g\|_q }[/math]

for all measurable real- or complex-valued functions f and g on S.

Notable special cases

For the following cases assume that p and q are in the open interval (1,∞) with 1/p + 1/q = 1.

Counting measure

For the n-dimensional Euclidean space, when the set S is {1, ..., n} with the counting measure, we have

[math]\displaystyle{ \sum_{k=1}^n |x_k\,y_k| \le \biggl( \sum_{k=1}^n |x_k|^p \biggr)^{\frac{1}{p}} \biggl( \sum_{k=1}^n |y_k|^q \biggr)^{\frac{1}{q}} \text{ for all }(x_1,\ldots,x_n),(y_1,\ldots,y_n)\in\mathbb{R}^n\text{ or }\mathbb{C}^n. }[/math]

Often the following practical form of this is used, for any [math]\displaystyle{ (r,s)\in\mathbb{R}_+ }[/math]:

[math]\displaystyle{ \biggl(\sum_{k=1}^n |x_k|^r\,|y_k|^s \biggr)^{r+s}\le \biggl( \sum_{k=1}^n |x_k|^{r+s} \biggr)^{r} \biggl( \sum_{k=1}^n |y_k|^{r+s} \biggr)^{s}. }[/math]

For more than two sums, the following generalisation ((Chen 2015)) holds, with real positive exponents [math]\displaystyle{ \lambda_i }[/math] and [math]\displaystyle{ \lambda_a + \lambda_b+ \cdots +\lambda_z =1 }[/math]:

[math]\displaystyle{ \sum_{k=1}^n |a_k|^{\lambda_a}\,|b_k|^{\lambda_b} \cdots |z_k|^{\lambda_z} \le \biggl(\sum_{k=1}^n |a_k|\biggr)^{\lambda_a} \biggl(\sum_{k=1}^n |b_k|\biggr)^{\lambda_b} \cdots \biggl(\sum_{k=1}^n |z_k|\biggr)^{\lambda_z} . }[/math]

Equality holds iff [math]\displaystyle{ |a_1|: |a_2|: \cdots : |a_n| =|b_1|: |b_2|: \cdots : |b_n| = \cdots = |z_1|: |z_2|: \cdots : |z_n| }[/math].

If S = N with the counting measure, then we get Hölder's inequality for sequence spaces:

[math]\displaystyle{ \sum_{k=1}^{\infty} |x_k\,y_k| \le \biggl( \sum_{k=1}^{\infty} |x_k|^p \biggr)^{\frac{1}{p}} \left( \sum_{k=1}^{\infty} |y_k|^q \right)^{\frac{1}{q}} \text{ for all }(x_k)_{k\in\mathbb N}, (y_k)_{k\in\mathbb N}\in\mathbb{R}^{\mathbb N}\text{ or }\mathbb{C}^{\mathbb N}. }[/math]

Lebesgue measure

If S is a measurable subset of Rn with the Lebesgue measure, and f and g are measurable real- or complex-valued functions on S, then Hölder inequality is

[math]\displaystyle{ \int_S \bigl| f(x)g(x)\bigr| \,\mathrm{d}x \le\biggl(\int_S |f(x)|^p\,\mathrm{d}x\biggr)^{\frac{1}{p}} \biggl(\int_S |g(x)|^q\,\mathrm{d}x\biggr)^{\frac{1}{q}}. }[/math]

Probability measure

For the probability space [math]\displaystyle{ (\Omega, \mathcal{F}, \mathbb{P}), }[/math] let [math]\displaystyle{ \mathbb{E} }[/math] denote the expectation operator. For real- or complex-valued random variables [math]\displaystyle{ X }[/math] and [math]\displaystyle{ Y }[/math] on [math]\displaystyle{ \Omega, }[/math] Hölder's inequality reads

[math]\displaystyle{ \mathbb{E}[|XY|] \leqslant \left (\mathbb{E}\bigl[ |X|^p\bigr]\right)^{\frac{1}{p}} \left(\mathbb{E}\bigl[|Y|^q\bigr]\right)^{\frac{1}{q}}. }[/math]

Let [math]\displaystyle{ 1 \lt r \lt s \lt \infty }[/math] and define [math]\displaystyle{ p = \tfrac{s}{r}. }[/math] Then [math]\displaystyle{ q = \tfrac{p}{p-1} }[/math] is the Hölder conjugate of [math]\displaystyle{ p. }[/math] Applying Hölder's inequality to the random variables [math]\displaystyle{ |X|^r }[/math] and [math]\displaystyle{ 1_{\Omega} }[/math] we obtain

[math]\displaystyle{ \mathbb{E}\bigl[|X|^r\bigr]\leqslant \left(\mathbb{E}\bigl[|X|^s\bigr]\right)^{\frac{r}{s}}. }[/math]

In particular, if the sth absolute moment is finite, then the r th absolute moment is finite, too. (This also follows from Jensen's inequality.)

Product measure

For two σ-finite measure spaces (S1, Σ1, μ1) and (S2, Σ2, μ2) define the product measure space by

[math]\displaystyle{ S=S_1\times S_2,\quad \Sigma=\Sigma_1\otimes\Sigma_2,\quad \mu=\mu_1\otimes\mu_2, }[/math]

where S is the Cartesian product of S1 and S2, the σ-algebra Σ arises as product σ-algebra of Σ1 and Σ2, and μ denotes the product measure of μ1 and μ2. Then Tonelli's theorem allows us to rewrite Hölder's inequality using iterated integrals: If f and g are Σ-measurable real- or complex-valued functions on the Cartesian product S, then

[math]\displaystyle{ \int_{S_1}\int_{S_2}|f(x,y)\,g(x,y)|\,\mu_2(\mathrm{d}y)\,\mu_1(\mathrm{d}x) \le\left(\int_{S_1}\int_{S_2}|f(x,y)|^p\,\mu_2(\mathrm{d}y)\,\mu_1(\mathrm{d}x)\right)^{\frac{1}{p}}\left(\int_{S_1}\int_{S_2}|g(x,y)|^q\,\mu_2(\mathrm{d}y)\,\mu_1(\mathrm{d}x)\right)^{\frac{1}{q}}. }[/math]

This can be generalized to more than two σ-finite measure spaces.

Vector-valued functions

Let (S, Σ, μ) denote a σ-finite measure space and suppose that f = (f1, ..., fn) and g = (g1, ..., gn) are Σ-measurable functions on S, taking values in the n-dimensional real- or complex Euclidean space. By taking the product with the counting measure on {1, ..., n}, we can rewrite the above product measure version of Hölder's inequality in the form

[math]\displaystyle{ \int_S \sum_{k=1}^n|f_k(x)\,g_k(x)|\,\mu(\mathrm{d}x) \le \left(\int_S\sum_{k=1}^n|f_k(x)|^p\,\mu(\mathrm{d}x)\right)^{\frac{1}{p}}\left(\int_S\sum_{k=1}^n|g_k(x)|^q\,\mu(\mathrm{d}x)\right)^{\frac{1}{q}}. }[/math]

If the two integrals on the right-hand side are finite, then equality holds if and only if there exist real numbers α, β ≥ 0, not both of them zero, such that

[math]\displaystyle{ \alpha \left (|f_1(x)|^p,\ldots,|f_n(x)|^p \right )= \beta \left (|g_1(x)|^q,\ldots,|g_n(x)|^q \right ), }[/math]

for μ-almost all x in S.

This finite-dimensional version generalizes to functions f and g taking values in a normed space which could be for example a sequence space or an inner product space.

Proof of Hölder's inequality

There are several proofs of Hölder's inequality; the main idea in the following is Young's inequality for products.

Alternative proof using Jensen's inequality:

Extremal equality

Statement

Assume that 1 ≤ p < ∞ and let q denote the Hölder conjugate. Then for every fLp(μ),

[math]\displaystyle{ \|f\|_p = \max \left \{ \left| \int_S f g \, \mathrm{d}\mu \right | : g\in L^q(\mu), \|g\|_q \le 1 \right\}, }[/math]

where max indicates that there actually is a g maximizing the right-hand side. When p = ∞ and if each set A in the σ-field Σ with μ(A) = ∞ contains a subset B ∈ Σ with 0 < μ(B) < ∞ (which is true in particular when μ is σ-finite), then

[math]\displaystyle{ \|f\|_\infty = \sup \left\{ \left| \int_S f g \,\mathrm{d}\mu \right| : g\in L^1(\mu), \|g\|_1 \le 1 \right \}. }[/math]

Proof of the extremal equality:

Remarks and examples

  • The equality for [math]\displaystyle{ p = \infty }[/math] fails whenever there exists a set [math]\displaystyle{ A }[/math] of infinite measure in the [math]\displaystyle{ \sigma }[/math]-field [math]\displaystyle{ \Sigma }[/math] with that has no subset [math]\displaystyle{ B \in \Sigma }[/math] that satisfies: [math]\displaystyle{ 0 \lt \mu(B) \lt \infty. }[/math] (the simplest example is the [math]\displaystyle{ \sigma }[/math]-field [math]\displaystyle{ \Sigma }[/math] containing just the empty set and [math]\displaystyle{ S, }[/math] and the measure [math]\displaystyle{ \mu }[/math] with [math]\displaystyle{ \mu(S) = \infty. }[/math]) Then the indicator function [math]\displaystyle{ 1_A }[/math] satisfies [math]\displaystyle{ \|1_A\|_{\infty} = 1, }[/math] but every [math]\displaystyle{ g \in L^1 (\mu) }[/math] has to be [math]\displaystyle{ \mu }[/math]-almost everywhere constant on [math]\displaystyle{ A, }[/math] because it is [math]\displaystyle{ \Sigma }[/math]-measurable, and this constant has to be zero, because [math]\displaystyle{ g }[/math] is [math]\displaystyle{ \mu }[/math]-integrable. Therefore, the above supremum for the indicator function [math]\displaystyle{ 1_A }[/math] is zero and the extremal equality fails.
  • For [math]\displaystyle{ p = \infty, }[/math] the supremum is in general not attained. As an example, let [math]\displaystyle{ S = \mathbb{N}, \Sigma = \mathcal{P}(\mathbb{N}) }[/math] and [math]\displaystyle{ \mu }[/math] the counting measure. Define:
[math]\displaystyle{ \begin{cases} f: \mathbb{N} \to \mathbb{R} \\ f(n) = \frac{n-1}{n} \end{cases} }[/math]
Then [math]\displaystyle{ \|f\|_{\infty} = 1. }[/math] For [math]\displaystyle{ g \in L^1 (\mu, \mathbb{N}) }[/math] with [math]\displaystyle{ 0 \lt \|g\|_1 \leqslant 1, }[/math] let [math]\displaystyle{ m }[/math] denote the smallest natural number with [math]\displaystyle{ g(m) \neq 0. }[/math] Then
[math]\displaystyle{ \left |\int_S fg\,\mathrm{d}\mu\right| \leqslant \frac{m-1}{m}|g(m)|+\sum_{n=m+1}^\infty|g(n)| = \|g\|_1-\frac{|g(m)|}m\lt 1. }[/math]

Applications

  • The extremal equality is one of the ways for proving the triangle inequality ||f1 + f2||p ≤ ||f1||p + ||f2||p for all f1 and f2 in Lp(μ), see Minkowski inequality.
  • Hölder's inequality implies that every fLp(μ) defines a bounded (or continuous) linear functional κf on Lq(μ) by the formula
[math]\displaystyle{ \kappa_f(g) = \int_S f g \, \mathrm{d}\mu,\qquad g\in L^q(\mu). }[/math]
The extremal equality (when true) shows that the norm of this functional κf as element of the continuous dual space Lq(μ)* coincides with the norm of f in Lp(μ) (see also the Lp-space article).

Generalization with more than two functions

Statement

Assume that r (0, ∞] and p1, ..., pn (0, ∞] such that

[math]\displaystyle{ \sum_{k=1}^n \frac1{p_k} = \frac1r }[/math]

where 1/∞ is interpreted as 0 in this equation. Then for all measurable real or complex-valued functions f1, ..., fn defined on S,

[math]\displaystyle{ \left\|\prod_{k=1}^n f_k\right\|_r \le \prod_{k=1}^n \left\|f_k\right\|_{p_k} }[/math]

where we interpret any product with a factor of ∞ as ∞ if all factors are positive, but the product is 0 if any factor is 0.

In particular, if [math]\displaystyle{ f_k \in L^{p_k}(\mu) }[/math] for all [math]\displaystyle{ k \in \{ 1, \ldots, n \} }[/math] then [math]\displaystyle{ \prod_{k=1}^n f_k \in L^r(\mu). }[/math]

Note: For [math]\displaystyle{ r \in (0, 1), }[/math] contrary to the notation, ||.||r is in general not a norm because it doesn't satisfy the triangle inequality.

Proof of the generalization:

Interpolation

Let p1, ..., pn (0, ∞] and let θ1, ..., θn ∈ (0, 1) denote weights with θ1 + ... + θn = 1. Define [math]\displaystyle{ p }[/math] as the weighted harmonic mean, that is,

[math]\displaystyle{ \frac1p = \sum_{k=1}^n \frac{\theta_k}{p_k}. }[/math]

Given measurable real- or complex-valued functions [math]\displaystyle{ f_k }[/math] on S, then the above generalization of Hölder's inequality gives

[math]\displaystyle{ \left\| |f_1|^{\theta_1}\cdots |f_n|^{\theta_n}\right\|_p \le \left\||f_1|^{\theta_1}\right\|_{\frac{p_1}{\theta_1}}\cdots \left\| |f_n|^{\theta_n}\right\|_{\frac{p_n}{\theta_n}} = \|f_1\|_{p_1}^{\theta_1}\cdots \|f_n\|_{p_n}^{\theta_n}. }[/math]

In particular, taking [math]\displaystyle{ f_1 = \cdots = f_n=:f }[/math] gives

[math]\displaystyle{ \|f\|_p \leqslant \prod_{k=1}^n \|f\|_{p_k}^{\theta_k}. }[/math]

Specifying further θ1 = θ and θ2 = 1-θ, in the case [math]\displaystyle{ n = 2, }[/math] we obtain the interpolation result

Littlewood's inequality — For [math]\displaystyle{ \theta\in(0,1) }[/math] and [math]\displaystyle{ \frac {1}{p_\theta} = \frac{\theta}{p_1} + \frac {1-\theta}{p_0} }[/math],

[math]\displaystyle{ \|f\|_{p_\theta} \leqslant \|f\|_{p_1}^\theta \cdot \|f\|_{p_0}^{1-\theta}, }[/math]

An application of Hölder gives

Lyapunov's inequality — If [math]\displaystyle{ p = (1-\theta) p_0 + \theta p_1, \qquad \theta \in (0, 1), }[/math] then

[math]\displaystyle{ \left\| |f_0|^{\frac{p_0(1-\theta)}{p}} \cdot |f_1|^{\frac{p_1 \theta}{p}}\right\|_p^p \le \|f_0\|_{p_0}^{p_0(1-\theta)} \|f_1\|_{p_1}^{p_1\theta} }[/math]

and in particular

[math]\displaystyle{ \|f\|_p^p \leqslant \|f\|_{p_0}^{p_0(1-\theta)} \cdot \|f\|_{p_1}^{p_1\theta}. }[/math]

Both Littlewood and Lyapunov imply that if [math]\displaystyle{ f \in L^{p_0}\cap L^{p_1} }[/math] then [math]\displaystyle{ f \in L^p }[/math] for all [math]\displaystyle{ p_0 \lt p \lt p_1. }[/math][4]

Reverse Hölder inequalities

Two functions

Assume that p ∈ (1, ∞) and that the measure space (S, Σ, μ) satisfies μ(S) > 0. Then for all measurable real- or complex-valued functions f and g on S such that g(s) ≠ 0 for μ-almost all sS,

[math]\displaystyle{ \|fg\|_1\geqslant \|f\|_{\frac{1}{p}}\,\|g\|_{\frac{-1}{p-1}}. }[/math]

If

[math]\displaystyle{ \|fg\|_1 \lt \infty \quad \text{and} \quad \|g\|_{\frac{-1}{p-1}} \gt 0, }[/math]

then the reverse Hölder inequality is an equality if and only if

[math]\displaystyle{ \exists \alpha \geqslant 0 \quad |f| = \alpha|g|^{\frac{-p}{p-1}} \qquad \mu\text{-almost everywhere}. }[/math]

Note: The expressions:

[math]\displaystyle{ \|f\|_{\frac{1}{p}} }[/math] and [math]\displaystyle{ \|g\|_{\frac{-1}{p-1}}, }[/math]

are not norms, they are just compact notations for

[math]\displaystyle{ \left (\int_S|f|^{\frac{1}{p}}\,\mathrm{d}\mu\right)^{p} \quad \text{and} \quad \left (\int_S|g|^{\frac{-1}{p-1}}\,\mathrm{d}\mu\right)^{-(p-1)}. }[/math]
Proof of the reverse Hölder inequality (hidden, click show to reveal.)

Note that p and

[math]\displaystyle{ q:=\frac{p}{p-1}\in(1,\infty) }[/math]

are Hölder conjugates. Application of Hölder's inequality gives

[math]\displaystyle{ \begin{align} \left \||f|^{\frac{1}{p}}\right \|_1 &= \left \||fg|^{\frac{1}{p}}\,|g|^{-\frac{1}{p}}\right \|_1\\ &\leqslant \left \| |fg|^{\frac{1}{p}} \right \|_p \left \| |g|^{-\frac{1}{p}}\right \|_q \\ &=\|fg\|_1^{\frac{1}{p}}\left \||g|^{\frac{-1}{p-1}}\right \|_1^{\frac{p-1}{p}} \end{align} }[/math]

Raising to the power p gives us:

[math]\displaystyle{ \left \||f|^{\frac{1}{p}}\right \|_1^p \leqslant \|fg\|_1 \left \||g|^{\frac{-1}{p-1}}\right \|_1^{p-1}. }[/math]

Therefore:

[math]\displaystyle{ \left \||f|^{\frac{1}{p}}\right \|_1^p \left \||g|^{\frac{-1}{p-1}}\right \|_1^{-(p-1)} \leqslant \|fg\|_1 . }[/math]

Now we just need to recall our notation.

Since g is not almost everywhere equal to the zero function, we can have equality if and only if there exists a constant α ≥ 0 such that |fg| = α |g|q/p almost everywhere. Solving for the absolute value of f gives the claim.

Multiple functions

The Reverse Hölder inequality (above) can be generalized to the case of multiple functions if all but one conjugate is negative. That is,

Let [math]\displaystyle{ p_1,..., p_{m-1} \lt 0 }[/math] and [math]\displaystyle{ p_m \in \mathbb{R} }[/math] be such that [math]\displaystyle{ \sum_{k=1}^{m} \frac{1}{p_k} = 1 }[/math] (hence [math]\displaystyle{ 0 \lt p_m \lt 1 }[/math]). Let [math]\displaystyle{ f_k }[/math] be measurable functions for [math]\displaystyle{ k = 1,...,m }[/math]. Then
[math]\displaystyle{ \left\|\prod_{k=1}^n f_k\right\|_1 \ge \prod_{k=1}^n \left\|f_k\right\|_{p_k}. }[/math]

This follows from the symmetric form of the Hölder inequality (see below).

Symmetric forms of Hölder inequality

It was observed by Aczél and Beckenbach[5] that Hölder's inequality can be put in a more symmetric form, at the price of introducing an extra vector (or function):

Let [math]\displaystyle{ f = (f(1),\dots,f(m)) , g = (g(1),\dots, g(m)), h = (h(1),\dots,h(m)) }[/math] be vectors with positive entries and such that [math]\displaystyle{ f(i) g(i) h(i) = 1 }[/math] for all [math]\displaystyle{ i }[/math]. If [math]\displaystyle{ p,q,r }[/math] are nonzero real numbers such that [math]\displaystyle{ \frac{1}{p}+\frac{1}{q}+\frac{1}{r}=0 }[/math], then:

  • [math]\displaystyle{ \|f\|_p \|g\|_q \|h\|_r \ge 1 }[/math] if all but one of [math]\displaystyle{ p,q,r }[/math] are positive;
  • [math]\displaystyle{ \|f\|_p \|g\|_q \|h\|_r \le 1 }[/math] if all but one of [math]\displaystyle{ p,q,r }[/math] are negative.

The standard Hölder inequality follows immediately from this symmetric form (and in fact is easily seen to be equivalent to it). The symmetric statement also implies the reverse Hölder inequality (see above).

The result can be extended to multiple vectors:

Let [math]\displaystyle{ f_1, \dots, f_n }[/math] be [math]\displaystyle{ n }[/math] vectors in [math]\displaystyle{ \mathbb{R}^m }[/math] with positive entries and such that [math]\displaystyle{ f_1(i) \dots f_n(i) = 1 }[/math] for all [math]\displaystyle{ i }[/math]. If [math]\displaystyle{ p_1,\dots,p_n }[/math] are nonzero real numbers such that [math]\displaystyle{ \frac{1}{p_1}+\dots+\frac{1}{p_n}=0 }[/math], then:

  • [math]\displaystyle{ \|f_1\|_{p_1} \dots \|f_n\|_{p_n} \ge 1 }[/math] if all but one of the numbers [math]\displaystyle{ p_i }[/math] are positive;
  • [math]\displaystyle{ \|f_1\|_{p_1} \dots \|f_n\|_{p_n} \le 1 }[/math] if all but one of the numbers [math]\displaystyle{ p_i }[/math] are negative.

As in the standard Hölder inequalities, there are corresponding statements for infinite sums and integrals.

Conditional Hölder inequality

Let (Ω, F, [math]\displaystyle{ \mathbb{P} }[/math]) be a probability space, GF a sub-σ-algebra, and p, q (1, ∞) Hölder conjugates, meaning that 1/p + 1/q = 1. Then for all real- or complex-valued random variables X and Y on Ω,

[math]\displaystyle{ \mathbb{E}\bigl[|XY|\big|\,\mathcal{G}\bigr] \le \bigl(\mathbb{E}\bigl[|X|^p\big|\,\mathcal{G}\bigr]\bigr)^{\frac{1}{p}} \,\bigl(\mathbb{E}\bigl[|Y|^q\big|\,\mathcal{G}\bigr]\bigr)^{\frac{1}{q}} \qquad\mathbb{P}\text{-almost surely.} }[/math]

Remarks:

[math]\displaystyle{ \mathbb{E}[Z|\mathcal{G}] = \sup_{n\in\mathbb{N}}\,\mathbb{E}[\min\{Z,n\}|\mathcal{G}]\quad\text{a.s.} }[/math]
  • On the right-hand side of the conditional Hölder inequality, 0 times ∞ as well as ∞ times 0 means 0. Multiplying a > 0 with ∞ gives ∞.

Proof of the conditional Hölder inequality:

Hölder's inequality for increasing seminorms

Let S be a set and let [math]\displaystyle{ F(S, \mathbb{C}) }[/math] be the space of all complex-valued functions on S. Let N be an increasing seminorm on [math]\displaystyle{ F(S, \mathbb{C}), }[/math] meaning that, for all real-valued functions [math]\displaystyle{ f, g \in F(S, \mathbb{C}) }[/math] we have the following implication (the seminorm is also allowed to attain the value ∞):

[math]\displaystyle{ \forall s \in S \quad f(s) \geqslant g(s) \geqslant 0 \qquad \Rightarrow \qquad N(f) \geqslant N(g). }[/math]

Then:

[math]\displaystyle{ \forall f, g \in F(S, \mathbb{C}) \qquad N(|fg|) \leqslant \bigl(N(|f|^p)\bigr)^{\frac{1}{p}} \bigl(N(|g|^q)\bigr)^{\frac{1}{q}}, }[/math]

where the numbers [math]\displaystyle{ p }[/math] and [math]\displaystyle{ q }[/math] are Hölder conjugates.[6]

Remark: If (S, Σ, μ) is a measure space and [math]\displaystyle{ N(f) }[/math] is the upper Lebesgue integral of [math]\displaystyle{ |f| }[/math] then the restriction of N to all Σ-measurable functions gives the usual version of Hölder's inequality.


Distances based on Hölder inequality

Hölder inequality can be used to define statistical dissimilarity measures[7] between probability distributions. Those Hölder divergences are projective: They do not depend on the normalization factor of densities.

See also

Citations

  1. Roman 2008, p. 303 §12
  2. Maligranda, Lech (1998), "Why Hölder's inequality should be called Rogers' inequality", Mathematical Inequalities & Applications 1 (1): 69–83, doi:10.7153/mia-01-05 
  3. Guessab, A.; Schmeisser, G. (2013), "Necessary and sufficient conditions for the validity of Jensen's inequality", Archiv der Mathematik 100 (6): 561–570, doi:10.1007/s00013-013-0522-3, "under the additional assumption that [math]\displaystyle{ \varphi'' }[/math] exists, this inequality was already obtained by Hölder in 1889" 
  4. Wojtaszczyk, P. (1991). Banach Spaces for Analysts. Cambridge Studies in Advanced Mathematics. Cambridge: Cambridge University Press. ISBN 978-0-521-56675-9. https://www.cambridge.org/core/books/banach-spaces-for-analysts/7647AD5BB01A5ABBC96789F74DDE8A2F. 
  5. Beckenbach, E. F. (1980). General inequalities 2. International Series of Numerical Mathematics / Internationale Schriftenreihe zur Numerischen Mathematik / Série Internationale d'Analyse Numérique. 47. Birkhäuser Basel. pp. 145–150. doi:10.1007/978-3-0348-6324-7. ISBN 978-3-7643-1056-1. https://doi.org/10.1007/978-3-0348-6324-7. 
  6. For a proof see (Trèves 1967).
  7. Nielsen, Frank; Sun, Ke; Marchand-Maillet, Stephane (2017). "On Hölder projective divergences". Entropy 3 (19): 122. doi:10.3390/e19030122. Bibcode2017Entrp..19..122N. 

References

External links

| last=Tisdell
| first=Chris
| title=Holder's Inequality
| publisher=YouTube
| url=https://www.youtube.com/watch?v=kxQiKaIuyOg
| year=2012