1 Introduction

Classical complex interpolation of Banach spaces, due to Calderón [5] (see [3] and, for more recent developments, [7]) is based on constructing holomorphic hulls generated by certain families of holomorphic mappings. A slightly different approach proposed in [8] rests on plurisubharmonic geodesics. The notion has been originally considered, starting from 1987, for metrics on compact Kähler manifolds (see [10] and the bibliography therein), while its local counterpart for plurisubharmonic functions on bounded hyperconvex domains of \({{\mathbb {C}}}^n\) was introduced more recently in [4] and [18], see also [1].

In the simplest case, the geodesics we need can be described as follows. Denote by \(A=\{\zeta \in {{\mathbb {C}}}:\,0< \log |\zeta | < 1\}\) the annulus bounded by the circles \(A_j=\{\zeta :\, \log |\zeta |=j\}\), \(j=0,1\). Let \(\varOmega \) be a bounded hyperconvex domain in \({{\mathbb {C}}}^n\). Given two plurisubharmonic functions \(u_0,u_1\) in \(\varOmega \), equal to zero on \(\partial \varOmega \), we consider the class W of all plurisubharmonic functions \(u(z,\zeta )\) in \(\varOmega \times A\), such that

$$\begin{aligned} \limsup _{\zeta \rightarrow A_j} u(z,\zeta )\le u_j(z)\quad \forall z\in \varOmega . \end{aligned}$$

Its Perron envelope \({{\mathcal {P}}}_W(z,\zeta )=\sup \{u(z,\zeta ):\, u\in W\}\) belongs to the class and satisfies \({{\mathcal {P}}}_W(z,\zeta )={{\mathcal {P}}}_W(z,|\zeta |)\), which gives rise to the functions

$$\begin{aligned} u_t(z):={{\mathcal {P}}}_W(z,e^t), \quad 0<t<1, \end{aligned}$$

the geodesic between \(u_0\) and \(u_1\). When the functions \(u_j\) are bounded, the geodesic \(u_t\) tends to \(u_j\) as \(t\rightarrow j\), uniformly on \(\varOmega \). One of the main properties of the geodesics is that they linearize the pluripotential energy functional

$$\begin{aligned} {{\mathcal {E}}}(u)=\int _\varOmega u(dd^c u)^n, \end{aligned}$$

which means

$$\begin{aligned} {{\mathcal {E}}}(u_t)=(1-t)\,{{\mathcal {E}}}(u_0)+t\,{{\mathcal {E}}}(u_1); \end{aligned}$$
(1)

see the details in [4] and [18].

In [18], this was adapted to the case when the endpoints \(u_j\) are relative extremal functions \(\omega _{K_j}\) of non-pluripolar compact sets \(K_0,K_1\subset \varOmega \); we recall that

$$\begin{aligned} \omega _K(z)=\omega _{K,\varOmega }(z)=\limsup _{y\rightarrow z}\, {{\mathcal {P}}}_{{{\mathcal {N}}}_K}(y), \end{aligned}$$

where \({{{\mathcal {N}}}_K}\) is the collection of all negative plurisubharmonic functions u in \(\varOmega \) with \(u|_K\le -1\), see [14]. Note that

$$\begin{aligned} {{\mathcal {E}}}(\omega _{K})=-{{\text {Cap}}\,}(K), \end{aligned}$$

where

$$\begin{aligned} {{\text {Cap}}\,}(K)={{\text {Cap}}\,}(K, \varOmega )=(dd^c\omega _K)^n(\varOmega )=(dd^c\omega _K)^n(K) \end{aligned}$$

is the Monge–Ampère capacity of K relative to \(\varOmega \).

If each \(K_j\) is polynomially convex (i.e., coincides with its polynomial hull), then the functions \(u_j=-1\) exactly on \(K_j\) are continuous on \({\overline{\varOmega }}\), and the geodesics \(u_t\in C({\overline{\varOmega }}\times [0,1])\). Let

$$\begin{aligned} K_t=\{z\in \varOmega :\, u_t(z)=-1\},\quad 0< t< 1; \end{aligned}$$
(2)

then (1) implies:

$$\begin{aligned} {{\text {Cap}}\,}(K_t)\le (1-t)\,{{\text {Cap}}\,}(K_0)+t\,{{\text {Cap}}\,}(K_1). \end{aligned}$$
(3)

As was shown in [19], the functions \(u_t\) in general are different from the relative extremal functions of \(K_t\). Moreover, if the sets \(K_j\) are Reinhardt (toric), then \(u_t =\omega _{K_t}\) for some \(t\in (0,1)\) only if \(K_0=K_1\), so an equality in (3) is never possible unless the geodesic degenerates to a point.

Furthermore, in the toric case, the capacities (with respect to the unit polydisk \({{\mathbb {D}}}^n\)) were proved in [8] to be not just convex functions of t, as is depicted in (3), but logarithmically convex:

$$\begin{aligned} {{\text {Cap}}\,}(K_t,{{\mathbb {D}}}^n)\le {{\text {Cap}}\,}(K_0,{{\mathbb {D}}}^n)^{1-t}\,{{\text {Cap}}\,}(K_1,{{\mathbb {D}}}^n)^t. \end{aligned}$$
(4)

This was done by representing the capacities, due to [2], as (co)volumes of certain sets in \({ {\mathbb {R}}}^n\) and applying convex geometry methods to an operation of copolar addition introduced in [19]. Furthermore, the sets \(K_t\) in the toric situation were shown to be the geometric means (multiplicative combinations) of \(K_j\), exactly as in the Calderón complex interpolation theory. And again, an equality in (4) is possible only if \(K_0=K_1\). It is worth mentioning that the volumes of \(K_t\) satisfy the opposite Brunn–Minkowski inequality [6]:

$$\begin{aligned} {{\text {Vol}}}(K_t)\ge {{\text {Vol}}}(K_0)^{1-t} {{\text {Vol}}}(K_1)^t. \end{aligned}$$

In this note, we apply the geodesic technique to weighted relative extremal functions

$$\begin{aligned} u_j^c=c_j\,\omega _{K_j},\quad c_j>0, \end{aligned}$$

the sets \(K_t\) being replaced with the sets \( K^c_t\) where the functions \(u_t^c\) attain their minimal values, \(-c_t\). The function \(t\mapsto c_t\) turns out to be convex; moreover, it is actually linear, \(c_t=(1-t)\,c_0+t\,c_1\), provided \(K_0\cap K_1\ne \emptyset \). With such an interpolation, one can have \(u_t^c=c_t\,\omega _{K^c_t,\varOmega }\) for a non-degenerate geodesic, in which case there is no loss in the transition from the functional \({{\mathcal {E}}}(u_t^c)\) to the capacity \({{\text {Cap}}\,}(K^c_t)\). And in any case, we establish the weighted inequality

$$\begin{aligned} c_t^{n+1}{{\text {Cap}}\,}(K^c_t)\le (1-t)\,c_0^{n+1}{{\text {Cap}}\,}(K_0)+t\,c_1^{n+1}{{\text {Cap}}\,}(K_1), \end{aligned}$$

which, for a smart choice of the constants \(c_j\), is stronger than (3) and even (in the toric case) than (4). In particular, it implies that the function

$$\begin{aligned} t\mapsto \left( {{\text {Cap}}\,}(K_t^c)\right) ^{-\frac{1}{n+1}} \end{aligned}$$

is concave.

In the toric setting of Reinhardt sets \(K_j\) in the unit polydisk, we show that the interpolating sets \(K_t^c\) actually are the geometric means, so they do not depend on the weights \(c_j\) and coincide with the sets \(K_t\) in the non-weighted interpolation; we do not know if the latter is true in the general, non-toric setting.

Finally, we transfer the above results on the capacities of sets in \({{\mathbb {C}}}^n\) to the realm of convex geometry, developing thus the Brunn–Minkowski theory for volumes of (co)convex sets in \({ {\mathbb {R}}}^n\) [8, 12, 19].

2 General Setting

Here, we consider the general case of \(u_j^c=c_j\,\omega _{K_j}\) with \(c_j>0\) and \(K_j\) non-pluripolar, compact, polynomially convex subsets of a bounded hyperconvex domain \(\varOmega \) of \({{\mathbb {C}}}^n\). In this situation, the functions \(u_j^c(z)=-c_j\) precisely on \(K_j\) and are continuous on \({\overline{\varOmega }}\), the geodesics \(u_t\) converge to \(u_j\), uniformly on \(\varOmega \), as \(t\rightarrow j\), and belong to \(C({\overline{\varOmega }}\times [0,1])\), as in the non-weighted case dealt with in [18] and [8].

Denote:

$$\begin{aligned} c_t=-\min \{u_t^c(z):\, z\in \varOmega \} \end{aligned}$$

and

$$\begin{aligned} K^c_t=\{z\in \varOmega :\, u^c_t(z)=-c_t\},\quad 0< t< 1, \end{aligned}$$
(5)

the set where \(u_t^c\) attains its minimal value on \(\varOmega \).

Theorem 1

In the above setting, we have:

  1. (i)

    \( c_t\le (1-t)\,c_0+t\,c_1\), with an equality if \(K_0\cap K_1\ne \emptyset \);

  2. (ii)

    the function \(t\mapsto c_t^{n+1}{{\text {Cap}}\,}(K_t)\) is convex:

    $$\begin{aligned} c_t^{n+1}{{\text {Cap}}\,}(K^c_t)\le (1-t)\,c_0^{n+1}{{\text {Cap}}\,}(K_0)+t\,c_1^{n+1}{{\text {Cap}}\,}(K_1); \end{aligned}$$
    (6)
  3. (iii)

    if the weights \(c_j\) are chosen such that

    $$\begin{aligned} c_0^{n+1}{{\text {Cap}}\,}(K_0)=c_1^{n+1}{{\text {Cap}}\,}(K_1), \end{aligned}$$
    (7)

    then the function

    $$\begin{aligned} V(t):=\left( {{\text {Cap}}\,}(K_t^c)\right) ^{-\frac{1}{n+1}} \end{aligned}$$

    is concave and, consequently, the function

    $$\begin{aligned} \rho (t):=V(t)^{-1}=\left( {{\text {Cap}}\,}(K_t^c)\right) ^{\frac{1}{n+1}} \end{aligned}$$

    is convex.

Proof

(i):

Consider \(v_j = c_j\,\omega _K\) for \(j=0,1\), where \(K=K_0\cup K_1\). The set K might be not polynomially convex, but \(\omega _K\) is nevertheless a bounded plurisubharmonic function on \(\varOmega \) with zero boundary values, so the geodesic \(v_t^c\) is well defined and converge to \(v_j\), uniformly on \(\varOmega \), as \(t\rightarrow j\) [18, Prop. 3.1]. Since \(v_j\le u_j^c\), we have \(v_t^c\le u_t^c\). Assume \(c_0\ge c_1\), then the corresponding geodesic \(v_t^c = \max \{c_0\,\omega _K, -((1-t)\,c_0+t\,c_1)\}\), because the right-hand side is maximal in \(\varOmega \times A\) and has the prescribed boundary values at \(t=0\) and \(t=1\). Therefore:

$$\begin{aligned} -c_t\ge \min \{v_t^c(z):\, z\in \varOmega \}\ge -((1-t)\,c_0+t\,c_1), \end{aligned}$$

which proves the convexity of \(c_t\).

To finish the proof of (i), let \(z^*\in K_0\cap K_1\ne \emptyset \), then \(-c_t\le u_t^c(z^*)\). Since the convexity of the function \(u_t^c(z^*)\) in t implies \(u_t^c(z^*)\le -((1-t)\,c_0+t\,c_1)\), we get \( c_t\ge (1-t)\,c_0+t\,c_1\) and thus the linearity.

(ii):

Since \(u_j^c=c_j\,\omega _{K_j}\), we have:

$$\begin{aligned} {{\mathcal {E}}}(u_j)= c_j^{n+1}\int _\varOmega (dd^c \omega _{K_j})^n = -c_j^{n+1}{{\text {Cap}}\,}(K_j), \quad j=1,2. \end{aligned}$$

For any fixed t, the function \(u_t^c=-c_t\) on \(K_t^c\), so \(u_t^c\le -c_t\,\omega _{K_t^c}\). By [18, Cor. 2.2] this implies

$$\begin{aligned} {{\mathcal {E}}}(u_t^c)\le {{\mathcal {E}}}(c_t\,\omega _{K_t^c})=-c_t^{n+1}{{\text {Cap}}\,}(K_t^c), \end{aligned}$$

and (6) follows from the geodesic linearization property (1).

(iii):

It suffices to prove the concavity of the function V. When the weights \(c_j\) satisfy (7), inequality (6) rewrites as

$$\begin{aligned} V(t)\ge \frac{c_t}{c_0}V(0) \end{aligned}$$

and, since

$$\begin{aligned} c_1= \frac{V(1)}{V(0)}c_0, \end{aligned}$$

this gives us

$$\begin{aligned} V(t)\ge (1-t)\,V(0)+t\,V(1), \end{aligned}$$

which completes the proof.

\(\square \)

The convexity/concavity results in this theorem are stronger than inequality (3) obtained in [18] by the geodesic interpolation \(u_t\) of non-weighted extremal functions. In addition, the non-weighted geodesic \(u_t\) is very unlikely to be the extremal function of the set \(K_t\) (as shown in [19], this is never the case in the toric situation, unless \(K_0=K_1\)), while this is perfectly possible in the weighted interpolation. For example, given \(K_0\Subset \varOmega \), let

$$\begin{aligned} K_1=\left\{ z\in \varOmega :\, \omega _{K_0}(z)\le -\frac{1}{2}\right\} , \end{aligned}$$

then \(\omega _{K_1,\varOmega }=\max \{2\omega _{K_0,\varOmega },-1\}\). For \(c_0=2\) and \(c_1=1\), we get:

$$\begin{aligned} u_t^c=\max \{2\omega _{K_0},-2+t\}= (2-t)\,\omega _{K_t^c} \end{aligned}$$

with

$$\begin{aligned} K_t^c=\{z\in \varOmega :\, \omega _{K_0}(z)\le -1+t/2\}, \end{aligned}$$

so

$$\begin{aligned} {{\text {Cap}}\,}(K_t^c)=\left( 1-\frac{t}{2}\right) ^{-1}|{{\mathcal {E}}}(u_t^c)|=\left( 1-\frac{t}{2}\right) ^{-1-n}{{\text {Cap}}\,}(K_0). \end{aligned}$$

3 Toric Case

In this section, we assume \(\varOmega ={{\mathbb {D}}}^n\), the unit polydisk, and \(K_0,K_1\subset {{\mathbb {D}}}^n\) to be non-pluripolar, polynomially convex compact Reinhardt (multicircled, toric) sets. Polynomial convexity of such a set K means that its logarithmic image

$$\begin{aligned} {{\text {Log}}\,}K=\{s\in {{\mathbb {R}}}_-^n:\, (e^{s_1},\ldots ,e^{s_n})\in K\} \end{aligned}$$

is a complete convex subset of \({{\mathbb {R}}}_-^n\), i.e., \({{\text {Log}}\,}K+{{\mathbb {R}}}_-^n\subset \log K\); we will also say that K is complete logarithmically convex. The functions \(c_j\,\omega _{K_j}\) are toric, and so is their geodesic \(u_t^c\). Note that since \(K_0\cap K_1\ne \emptyset \), inequality (6) and the concavity/convexity statements of Theorem 1(iii) hold true.

It was shown in [8] that the sets \(K_t\) defined by (2) for the geodesic interpolation of non-weighted toric extremal functions \(\omega _{K_j}\) are, as in the classical interpolation theory, the geometric means \(K_t^\times \) of \(K_j\). Here, we extend the result to the weighted interpolation, which shows, in particular, that the sets \(K_t^c\) do not depend on the weights \(c_j\). The relation \(K_t^\times \subset K_t^c\) is easy, while the reverse inclusion is more elaborate; we mimic the proof of the corresponding relation for the non-weighted case [8] that rests on a machinery developed in [19].

Any toric plurisubharmonic function u(z) in \({{\mathbb {D}}}^n\) gives rise to a convex function

$$\begin{aligned} {\check{u}}(s)=u (e^{s_1},\ldots ,e^{s_n}), \quad s\in {{\mathbb {R}}}_-^n, \end{aligned}$$
(8)

and the geodesic \(u_t^c\) generates the function \({\check{u}}_t^c\), convex in \((s,t)\in {{\mathbb {R}}}_-^n\times (0,1)\).

Given a convex function f on \({{\mathbb {R}}}_-^n\), we extend it to the whole \({ {\mathbb {R}}}^n\) as a lower semicontinuous convex function on \({ {\mathbb {R}}}^n\), equal to \(+\infty \) on \({ {\mathbb {R}}}^n{\setminus }{{\overline{{{\mathbb {R}}}_-^n}}}\), and we denote \({{\mathcal {L}}}[f]\) its Legendre transform:

$$\begin{aligned} {{\mathcal {L}}}[f](x)=\sup _{y\in { {\mathbb {R}}}^n}\{\langle x,y\rangle -f(y)\}. \end{aligned}$$

Evidently, \({{\mathcal {L}}}[f](x)=+\infty \) if \(x\not \in \overline{{{\mathbb {R}}}_+^n}\), and the Legendre transform is an involutive duality between convex functions on \({{\mathbb {R}}}_+^n\) and \({{\mathbb {R}}}_-^n\).

It was shown in [19] that for the relative extremal function \(\omega _K=\omega _{K,{{\mathbb {D}}}^n}\)

$$\begin{aligned} {{\mathcal {L}}}[{\check{\omega }}_K]=\max \{h_Q+1,0\}, \end{aligned}$$

where

$$\begin{aligned} h_Q(a)=\sup _{s\in Q} \langle a,s\rangle ,\quad a\in {{\mathbb {R}}}_+^n\end{aligned}$$

is the support function of the convex set \(Q={{\text {Log}}\,}K \subset {{\mathbb {R}}}_-^n\). Therefore, for a weighted relative extremal function \(u=c\,\omega _{K}\), we have:

$$\begin{aligned} {{\mathcal {L}}}[{\check{u}}](a) = c_j\,{{\mathcal {L}}}[{\check{\omega }}_{K_j}](c_j^{-1}a) = \max \{h_Q(a)+c_j,0\}. \end{aligned}$$
(9)

Theorem 2

Given two non-pluripolar complete logarithmically convex compact Reinhardt sets \(K_0,K_1\subset {{\mathbb {D}}}^n\) and two positive numbers \(c_0\) and \(c_1\), let \(u_t^c\) be the geodesic connecting the functions \(u_0=c_0\,\omega _{K_0}\) and \(u_1=c_1\,\omega _{K_1}\). Then the interpolating sets \(K_t^c\) defined by (5) coincide with the geometric means:

$$\begin{aligned} K_t^\times :=K_0^{1-t}K_1^t=\{z:\, |z_l|=|\eta _l|^{1-t} |\xi _l|^{t}, \ 1\le l\le n,\ \eta \in K_0,\ \xi \in K_1\}. \end{aligned}$$

Proof

Since the sets \(K_t^\times \) and \(K_t^c\) are complete logarithmically convex, it suffices to prove that \(Q_t:={{\text {Log}}\,}K_t^\times \) coincides with \(Q_t^c:={{\text {Log}}\,}K_t^c\).

The inclusion \(Q_t\subset Q_t^c\) follows from convexity of the function \({\check{u}}_t^c(s)\) in \((s,t)\in {{\mathbb {R}}}_-^n\times (0,1)\): if \(s\in Q_t\), then \(s=(1-t)\,s_0+t\,s_1\) for some \(s_j\in Q_j\), so:

$$\begin{aligned} {\check{u}}_t(s)\le (1-t)\, {\check{u}}_0(s_0)+ t\, {\check{u}}_1(s_1)=c_t, \end{aligned}$$

while we have \({\check{u}}_t(s)\ge -c_t\) for all s. This gives us \(s\in Q_t^c\).

To prove the reverse inclusion, take an arbitrary point \(\xi \in {{\mathbb {R}}}_-^n{\setminus } Q_t\), then there exists \(b\in {{\mathbb {R}}}_+^n\), such that

$$\begin{aligned} \langle b,\xi \rangle > h_{Q_t}(b)=(1-t)h_{Q_0}(b)+ t\, h_{Q_1}(b). \end{aligned}$$
(10)

By the homogeneity of the both sides, we can assume \(h_{Q_0}(b)\ge -c_0\) and \(h_{Q_1}(b)\ge -c_1\). Then, by (9) and (10), we have:

$$\begin{aligned} {\check{u}}_t(\xi )= & {} \sup _{a}[\langle a,\xi \rangle - (1-t)\max \{h_{Q_0}(a)+c_0,0\} - t \max \{h_{Q_1}(a)+c_1,0\}]\\\ge & {} \langle b,\xi \rangle - (1-t)\max \{h_{Q_0}(b)+c_0,0\} - t \max \{h_{Q_1}(b)+c_1,0\}\\> & {} (1-t)[h_{Q_0}(b)-(h_{Q_0}(b)+1)] + t [h_{Q_1}(b)-(h_{Q_1}(b)+1)]=-1, \end{aligned}$$

so \(\xi \not \in Q_t^c\). \(\square \)

Now, the corresponding assertions of Theorem 1 can be stated as follows.

Theorem 3

For non-pluripolar complete logarithmically convex compact Reinhardt sets \(K_0,K_1\subset {{\mathbb {D}}}^n\), the inequality

$$\begin{aligned} c_t^{n+1}{{\text {Cap}}\,}(K_t^\times , {{\mathbb {D}}}^n)\le (1-t)\,c_0^{n+1}{{\text {Cap}}\,}(K_0,{{\mathbb {D}}}^n)+t\,c_1^{n+1}{{\text {Cap}}\,}(K_1,{{\mathbb {D}}}^n) \end{aligned}$$
(11)

holds true for any \(c_0,c_1>0\) and \(c_t=(1-t)\,c_0+t\,c_1\).

In particular, the function

$$\begin{aligned} t\mapsto \left( {{\text {Cap}}\,}(K_t^\times ,{{\mathbb {D}}}^n)\right) ^{-\frac{1}{n+1}} \end{aligned}$$

is concave and consequently the function

$$\begin{aligned} t\mapsto \left( {{\text {Cap}}\,}(K_t^\times ,{{\mathbb {D}}}^n)\right) ^{\frac{1}{n+1}} \end{aligned}$$

is convex.

As we saw in the example in the previous section, sometimes one can have \(u_t=\omega _{K_t^c}\) for \(u_j=c_j\,\omega _{K_j}\), in which case (11) becomes an equality. Our next result determines when this is possible for the toric case.

Theorem 4

In the conditions of Theorem 2, the geodesic \(u_\tau ^c\) equals \(c_\tau \,\omega _{K_\tau }\) for some \(\tau \in (0,1)\) if and only if

$$\begin{aligned} K_1^{c_0}=K_0^{c_1}, \end{aligned}$$

that is, \(c_0\,{{\text {Log}}\,}K_1 =c_1\,{{\text {Log}}\,}K_0\).

Proof

We will use the toric geodesic representation formula established in [19, Thm. 5.1]:

$$\begin{aligned} {\check{u}}_t ={{\mathcal {L}}}\left[ (1-t){{\mathcal {L}}}[{\check{u}}_0] + t{{\mathcal {L}}}[{\check{u}}_1]\right] , \end{aligned}$$
(12)

which is a local counterpart of Guan’s result [9] for compact toric manifolds; here, \({\check{u}}\) is the convex image (8) of the toric plurisubharmonic function u.

Let \(Q_t=\log K_t\), \(0\le t\le 1\). By (9), \(u_\tau ^c=c_\tau \,\omega _{K_\tau }\) means

$$\begin{aligned} (1-\tau )\max \{h_{Q_0}(a)+c_0,0\} + \tau \max \{h_{Q_1}(a)+c_1,0\}=\max \{h_{Q_\tau }(a)+c_\tau ,0\}, \end{aligned}$$

or, which is the same,

$$\begin{aligned} \max \{h_{(1-\tau )Q_0}(a)+(1-\tau )c_0,0\} + \max \{h_{\tau Q_1}(a)+\tau c_1,0\}=\max \{h_{Q_\tau }(a)+c_\tau ,0\} \end{aligned}$$

for all \(a\in {{\mathbb {R}}}_+^n\). Therefore, \(h_{Q_0}(a)\le -c_0\) if and only if \(h_{Q_1}(a)\le -c_1\), so \(c_0\,Q_0^\circ = c_1\,Q_1^\circ \) and, since \((c \,Q)^\circ =c^{-1}Q^\circ \), we get \(c_0\,Q_1=c_1\,Q_0\). Here \(Q^\circ \) is the copolar (14) to the set Q, see the beginning of the next section. \(\square \)

4 Covolumes

In the toric case, the Monge–Ampère capacities with respect to the unit polydisk can be represented as volumes of certain sets [2, 19]. Namely, if \(K\Subset {{\mathbb {D}}}^n\) is complete and logarithmically convex, then \(Q:={{\text {Log}}\,}K\subset {{\mathbb {R}}}_-^n\) and

$$\begin{aligned} {{\text {Cap}}\,}(K,{{\mathbb {D}}}^n)=n!\,{{\text {Covol}}}(Q^\circ ):=n!\, {{\text {Vol}}}({{{\mathbb {R}}}_+^n{\setminus } Q^\circ }), \end{aligned}$$
(13)

where the convex set \(Q^\circ \subset {{\mathbb {R}}}_+^n\) defined by

$$\begin{aligned} Q^\circ =\{x\in { {\mathbb {R}}}^n: h_Q(x) \le -1 \}= \{x\in { {\mathbb {R}}}^n: \langle x,y\rangle \le -1 \ \forall y\in Q\} \end{aligned}$$
(14)

is, in the terminology of [19], the copolar to the set Q. In particular:

$$\begin{aligned} {{\text {Cap}}\,}(K_t^\times ,{{\mathbb {D}}}^n)=n!\,{{\text {Covol}}}(Q_t^\circ ) \end{aligned}$$

for the copolar \(Q_t^\circ \) of the set \(Q_t=(1-t)Q_0+t\,Q_1\); we would like to stress that \(Q_t^\circ \ne (1-t)Q_0^\circ +t\,Q_1^\circ \).

Convex complete subsets P of \({{\mathbb {R}}}_+^n\) (i.e., \(P+{{\mathbb {R}}}_+^n\subset P\)) appear in singularity theory and complex analysis (see, for example, [11,12,13, 15,16,17]), their covolumes (the volumes of \({{\mathbb {R}}}_+^n{\setminus } P\)) being used for computation of the multiplicities of mappings, etc. Such a set P generates, by the same formula (14), its copolar \(P^\circ \subset {{\mathbb {R}}}_-^n\), whose exponential image \({{\text {Exp}}\,}P^\circ \) (the closure of all points \((e^{s_1},\ldots ,e^{s_n})\) with \(s\in P^\circ \)) is a complete logarithmically convex subset of \({{\mathbb {D}}}^n\). Since taking the copolar is an involution, the representation (13) translates coherently the inequalities on the capacities to those on the (co)volumes. Namely, \({{\text {Cap}}\,}(Q_j)\) becomes replaced by \({{\text {Covol}}}(P_j)\) with \(P_j=Q_j^\circ \subset {{\mathbb {R}}}_+^n\) for \(j=0,1\), while \({{\text {Cap}}\,}(Q_t)\) has to be replaced with the covolume of the set whose copolar is \(Q_t\), that is, with \(\left( (1-t)\,P_0^\circ +t\,P_1^\circ \right) ^\circ \). The operation of copolar addition

$$\begin{aligned} P_0\oplus P_1:=\left( P_0^\circ +P_1^\circ \right) ^\circ \end{aligned}$$

was introduced in [19]. In particular, it was shown there that the copolar sum of any pair of cosimplices in \({{\mathbb {R}}}_+^n\), unlike their Minkowski sum, is still a simplex.

Corollary 1

Let \(P_0,P_1\) be non-empty convex complete subsets of \({{\mathbb {R}}}_+^n\), and let the interpolating sets \(P_t^\oplus \) be defined by

$$\begin{aligned} P_t^\oplus =\left( (1-t)P_0^\circ +tP_1^\circ \right) ^\circ ,\quad 0< t< 1. \end{aligned}$$

Then the inequality

$$\begin{aligned} c_t^{n+1}{{\text {Covol}}}(P_t^\oplus )\le (1-t)\,c_0^{n+1}{{\text {Covol}}}(P_0)+t\,c_1^{n+1}{{\text {Covol}}}(P_1) \end{aligned}$$

holds true for any \(c_0,c_1>0\) and \(c_t=(1-t)\,c_0+t\,c_1\).

In particular, the function

$$\begin{aligned} V^\oplus [P](t):=\left( {{\text {Covol}}}(P_t^\oplus )\right) ^{-\frac{1}{n+1}} \end{aligned}$$

is concave and, consequently, the function

$$\begin{aligned} \rho ^\oplus [P](t):=\left( {{\text {Covol}}}(P_t^\oplus )\right) ^{\frac{1}{n+1}} \end{aligned}$$

is convex.

Note that the convexity of \(\rho ^\oplus \) (following from the concavity of \(V^\oplus \)) implies that the function

$$\begin{aligned} {\tilde{\rho }}^\oplus [P](t):= \left( {{\text {Covol}}}(P_t^\oplus )\right) ^{\frac{1}{n}} \end{aligned}$$

is convex as well. Since \({\tilde{\rho }}^\oplus \) is a homogeneous function of P, that is,

$$\begin{aligned} {\tilde{\rho }}^\oplus [c\,P](t)=c\,{\tilde{\rho }}^\oplus [P](t) \end{aligned}$$

for all \(c>0\) and \(0<t<1\), its convexity is equivalent to the logarithmic convexity of the covolumes, established in [8] by convex geometry methods:

$$\begin{aligned} {{\text {Covol}}}(P_t^\oplus )\le {{\text {Covol}}}(P_0)^{1-t}{{\text {Covol}}}(P_1)^t, \end{aligned}$$

which is just another form of the Brunn–Minkowski type inequality (4). Therefore, the concavity of the function \(V^\oplus \) is a stronger property than just the logarithmic convexity of the covolumes.