The shift-invariant space \(V^\infty (g)\) enjoys the following sampling property [4]: every separated set \(\Lambda \subseteq {\mathbb {R}}\) with lower Beurling density

$$\begin{aligned} D^{-}(\Lambda ) := \liminf _{r \longrightarrow \infty } \inf _{x\in {\mathbb {R}}} \frac{\# \big [ \,\Lambda \cap [x-r,x+r] \,\big ]}{2r} \end{aligned}$$

strictly larger than 1 provides the norm equivalence

$$\begin{aligned} ||{f}||_{L^\infty ({\mathbb {R}})} \asymp ||{f|\Lambda }||_{\ell ^\infty ({\Lambda })}, \quad f \in V^\infty (g). \end{aligned}$$
(2)

(A similar property holds for all \(L^p\) norms, \(1 \le p \le \infty \); see [4, Theorems 5.2 and 3.1].)

In this article, we show the following uniqueness property for the absolute values of real-valued functions in \(V^\infty (g)\).

FormalPara Theorem 1

(Sign retrieval) Let g be a totally positive function of Gaussian type, as in (1), and \(\Lambda \subseteq {\mathbb {R}}\) with lower Beurling density

$$\begin{aligned} D^{-}(\Lambda )>2. \end{aligned}$$
(3)

Assume that \(f_1, f_2 \in V^\infty (g)\) are real-valued and \(\left| f_1 \right| \equiv \left| f_2 \right| \) on \(\Lambda \). Then either \(f_1 \equiv f_2\) or \(f_1 \equiv - f_2\).

See [2] for motivation for sign retrieval in shift-invariant spaces. When g is a Gaussian function—corresponding to \(m=0\) in (1)—Theorem 1 was recently obtained in [3, Theorem 1]. Here, that result is extended to all totally positive functions of Gaussian type.Footnote 1

The intuition behind Theorem 1 is as follows. Suppose that \(\left| f_1 \right| \equiv \left| f_2 \right| \) on \(\Lambda \), and split \(\Lambda \) into

$$\begin{aligned} \begin{aligned} \Lambda _1&= \{ \lambda \in \Lambda : f_1(\lambda ) = f_2(\lambda )\},\\ \Lambda _2&= \{ \lambda \in \Lambda : f_1(\lambda ) = -f_2(\lambda )\}. \end{aligned} \end{aligned}$$
(4)

Then \(f_1 - f_2\) vanishes on \(\Lambda _1\) while \(f_1 + f_2\) vanishes on \(\Lambda _2\). Under (3), one may expect one of the two subsets \(\Lambda _j\) to have lower Beurling density larger than 1. The sampling inequalities (2) would then imply that either \(f_1 - f_2\) or \(f_1 + f_2\) are identically zero. This argument breaks down, however, because Beurling’s lower density is not subadditive. For example, \(\Lambda _1 := \Lambda \cap (-\infty ,0]\) and \(\Lambda _2 := \Lambda \cap (0,\infty )\) have always zero lower Beurling density. The proof of the sign retrieval theorem for Gaussian generators in [3] resorts instead to a special property of the Gaussian function, namely that \(V^\infty (g) \cdot V^\infty (g)\) is contained in a dilation of \(V^\infty (\tilde{g})\) by a factor of 2, where \(\tilde{g}\) is another Gaussian function. Thus, in the Gaussian case, the sampling theorem can be applied after rescaling to the set \(\Lambda \) to conclude that \((f_1-f_2) \cdot (f_1+f_2) \equiv 0\), and, by analyticity, that either \(f_1 \equiv f_2\) or \(f_1 \equiv - f_2\). A similar argument applies to Paley–Wiener spaces [1, Theorem 2.5]. We are unaware of an analogous dilation property for general totally positive generators.

To prove Theorem 1 for all totally positive generators of Gaussian type we take a different route. We define the upper average circular density of a set \(\Lambda \subseteq {\mathbb {R}}\) as

$$\begin{aligned} D^+_{\mathrm {circ}}(\Lambda ) = \limsup _{r \longrightarrow \infty } \frac{4}{\pi r^2} \int _0^r \sum _{\lambda \in \Lambda \cap [-t,t]} \sqrt{t^2 - \lambda ^2} \, \frac{dt}{t}, \end{aligned}$$
(5)

with the convention that \(D^+_{\mathrm {circ}}(\Lambda )=\infty \), if \(\Lambda \) is uncountably infinite. The density is named circular because, in (5), each point \(\lambda \in [-t,t]\) is weighted with the measure of the largest vertical segment \(\{\lambda \} \times (-a,a)\) contained in the two-dimensional open disk \(B_t(0) \subseteq {\mathbb {R}}^2\).

The upper average circular density can be alternatively described as follows: as shown in Lemma 1 below, for any lattice \(\alpha {\mathbb {Z}}\), \(\alpha >0\),

$$\begin{aligned} D^+_{\mathrm {circ}}(\Lambda )= \limsup _{r \longrightarrow \infty } \frac{2\alpha }{\pi r^2} \int _0^r \# \big [ (\Lambda \setminus \{0\} \times \alpha {\mathbb {Z}}) \cap B_t(0)\big ] \,\frac{dt}{t}. \end{aligned}$$
(6)

From here, it follows easily that \(D^+_{\mathrm {circ}}\) dominates Beurling’s lower density:

$$\begin{aligned} D^+_{\mathrm {circ}}(\Lambda ) \ge D^{-}(\Lambda ); \end{aligned}$$
(7)

see Lemma 1 below. We call \(D^+_{\mathrm {circ}}\) an upper density because, due to the sublinearity of \(\limsup \),

$$\begin{aligned} D^+_{\mathrm {circ}}(\Lambda _1 \cup \Lambda _2) \le D^+_{\mathrm {circ}}(\Lambda _1) + D^+_{\mathrm {circ}}(\Lambda _2), \end{aligned}$$
(8)

for any two sets \(\Lambda _1, \Lambda _2 \subseteq {\mathbb {R}}\).

Below we prove the following uniqueness result formulated in terms of the upper average circular density of the zero set \(\{f=0\}\) of a function f (counted without multiplicities).

FormalPara Theorem 2

(Uniqueness theorem) Let g be a totally positive function of Gaussian type, as in (1). Let \(f \in V^\infty (g)\) be non-zero. Then \(D^+_{\mathrm {circ}}(\{f=0\}) \le 1\).

The uniqueness theorem (which applies also to non-real-valued functions) allows one to carry out the following natural proof of the sign retrieval theorem.

FormalPara Proof

(Proof of Theorem 1, assuming Theorem 2) Assume that \(D^{-}(\Lambda )>2\) and write \(\Lambda = \Lambda _1 \cup \Lambda _2\) as in (4). Then, by (7) and (8), either \(D^+_{\mathrm {circ}}(\Lambda _1) >1\) or \(D^+_{\mathrm {circ}}(\Lambda _2) >1\) (possibly both). In the first case, Theorem 2 shows that \(f_1 \equiv f_2\), while in the second, \(f_1 \equiv -f_2\).\(\square \)

As the proof shows, Theorem 1 remains valid if (3) is relaxed to \(D^+_{\mathrm {circ}}(\Lambda )>2\). Sign retrieval also holds under the same density condition for the shift-invariant spaces \(V^p(g)\) defined with respect to \(L^p\) norms, \(1 \le p \le \infty \), as these are contained in \(V^\infty (g)\).

Towards the proof of Theorem 2, we first prove (6) and (7).

FormalPara Lemma 1

Let \(\Lambda \subseteq {\mathbb {R}}\) and \(\alpha >0\). Then (6) and (7) hold true.

FormalPara Proof

We assume that \(\Lambda \) has no accumulation points, since, otherwise, both sides of (6) are infinite. Denote provisionally the right hand side of (6) by \({\tilde{D}}^+_{\mathrm {circ}}(\Lambda )\), and set \(\Lambda ' := \Lambda \setminus \{0\}\).

Step 1. Let \(\varepsilon \in (0,1)\), and \(C=C_{\alpha ,\varepsilon }>0\) a constant to be specified. We claim:

$$\begin{aligned}&D^+_{\mathrm {circ}}(\Lambda ) = \limsup _{r \longrightarrow \infty } \frac{4}{\pi r^2} \int _{C} ^r \sum _{\lambda \in \Lambda ' \cap [-t,t]} \sqrt{t^2 - \lambda ^2} \, \frac{dt}{t}, \end{aligned}$$
(9)
$$\begin{aligned}&(1+\delta )^{2} {\tilde{D}}^+_{\mathrm {circ}}(\Lambda ) = \limsup _{r \longrightarrow \infty } \frac{2\alpha }{\pi r^2} \int _{C}^r \# \big [ (\Lambda ' \times \alpha {\mathbb {Z}}) \cap B_{(1+\delta )t}(0) \big ] \,\frac{dt}{t}, \quad \delta =\pm \varepsilon . \end{aligned}$$
(10)

To prove the first claim, we first note that \(D^+_{\mathrm {circ}}(\Lambda )=D^+_{\mathrm {circ}}(\Lambda ')\), since, if \(0 \in \Lambda \), the contribution of the point \(\lambda =0\) is

$$\begin{aligned} \frac{4}{\pi r^2} \int _0^r dt = \frac{4}{\pi r}, \end{aligned}$$

and does not affect the limit on r. Similarly, since \(\sqrt{t^2 - \lambda ^2} \le t\),

$$\begin{aligned} \frac{4}{\pi r^2} \int _0^C \sum _{\lambda \in \Lambda ' \cap [-t,t]} \sqrt{t^2 - \lambda ^2} \frac{dt}{t} \le \# \big ( \Lambda ' \cap [-C,C]\big ) \cdot \frac{4C}{\pi r^2}. \end{aligned}$$

The last quantity is finite because \(\Lambda \) has no accumulation points. This proves (9), because it shows that limiting the integral to [Cr] does not affect the limit on r. Second, a change of variables shows that (10) holds with C replaced by 0. In addition, since \(\Lambda \) has no accumulation points, there exists \(\eta >0\) such that \(\Lambda '\cap (-\eta ,\eta )=\emptyset \). Therefore, we can estimate the part of the integral excluded in (10) as

$$\begin{aligned} \frac{2\alpha }{\pi r^2} \int _{0}^C \# \big [(\Lambda ' \times \alpha {\mathbb {Z}}) \cap B_{(1+\delta )t}(0) \big ]\,\frac{dt}{t} \le \# \big [ (\Lambda ' \times \alpha {\mathbb {Z}}) \cap B_{2C}(0) \big ] \cdot \frac{2\alpha }{\pi r^2} \int _{\frac{\eta }{1+\varepsilon }}^C \frac{dt}{t}, \end{aligned}$$

which is finite because \(\eta >0\). This proves (10).

Step 2. For \(t>0\), we note that

$$\begin{aligned} \# (\Lambda ' \times \alpha {\mathbb {Z}}) \cap B_t(0)&= \sum _{\lambda \in \Lambda ' \cap [-t,t]} \# \left\{ k \in {\mathbb {Z}}: \lambda ^2+\alpha ^2 k^2< t^2 \right\} \\&= \sum _{\lambda \in \Lambda ' \cap [-t,t]} \# \left\{ k \in {\mathbb {Z}}: \left| k \right| < \alpha ^{-1} \sqrt{t^2 - \lambda ^2} \right\} . \end{aligned}$$

Next, we choose the constant \(C_{\alpha , \varepsilon }\) so that the following estimates hold for \(t \ge C_{\alpha , \varepsilon }\):

$$\begin{aligned}&\# \big [ (\Lambda ' \times \alpha {\mathbb {Z}}) \cap B_{(1+\varepsilon )t}(0) \big ] \ge \sum _{\lambda \in \Lambda ' \cap [-t,t]} \left[ \frac{2}{\alpha } \sqrt{(1+\varepsilon )^2 t^2 - \lambda ^2} - 1\right] \\&\quad =\sum _{\lambda \in \Lambda ' \cap [-t,t]} \frac{2}{\alpha } \sqrt{(1+\varepsilon )^2 t^2 - \lambda ^2 - \alpha \sqrt{(1+\varepsilon )^2 t^2 - \lambda ^2} + \alpha ^2/4 }\\&\quad \ge \sum _{\lambda \in \Lambda ' \cap [-t,t]} \frac{2}{\alpha } \sqrt{(1+\varepsilon )^2 t^2 - \lambda ^2 - \alpha (1+\varepsilon )t + \alpha ^2/4 }\\&\quad \ge \sum _{\lambda \in \Lambda ' \cap [-t,t]} \frac{2}{\alpha } \sqrt{t^2 - \lambda ^2}, \end{aligned}$$

where the last estimate requires choosing \(C_{\alpha , \varepsilon }\) large. Similarly,

$$\begin{aligned}&\# \big [(\Lambda ' \times \alpha {\mathbb {Z}}) \cap B_{(1-\varepsilon )t}(0)\big ] \le \sum _{\lambda \in \Lambda ' \cap [-(1-\varepsilon )t,(1-\varepsilon )t]} \left[ \frac{2}{\alpha } \sqrt{(1-\varepsilon )^2 t^2 - \lambda ^2} + 1\right] \\&\quad = \sum _{\lambda \in \Lambda ' \cap [-(1-\varepsilon )t,(1-\varepsilon )t]} \frac{2}{\alpha } \sqrt{(1-\varepsilon )^2 t^2 - \lambda ^2 + \alpha \sqrt{(1-\varepsilon )^2 t^2 - \lambda ^2} +\alpha ^2/4 } \\&\quad \le \sum _{\lambda \in \Lambda ' \cap [-(1-\varepsilon )t,(1-\varepsilon )t]} \frac{2}{\alpha } \sqrt{(1-\varepsilon )^2 t^2 - \lambda ^2 + \alpha {(1-\varepsilon ) t} +\alpha ^2/4 } \\&\quad \le \sum _{\lambda \in \Lambda ' \cap [-t,t]} \frac{2}{\alpha } \sqrt{t^2 - \lambda ^2}, \end{aligned}$$

where the last estimate requires choosing \(C_{\alpha , \varepsilon }\) large. Combining the last two estimates with (9) and (10), and letting \(\varepsilon \longrightarrow 0\), we deduce (6).

Step 3. To prove (7), let \(r>s>0\) and use (6) with \(\alpha =1\) to estimate

$$\begin{aligned} D^+_{\mathrm {circ}}(\Lambda )&\ge \liminf _{r \longrightarrow \infty } \frac{2}{\pi r^2} \int _s^r \inf _{z \in {\mathbb {R}}^2} \# \big [ (\Lambda ' \times {\mathbb {Z}}) \cap B_t(z) \big ] \,\frac{dt}{t} \\&\ge \inf _{t \ge s} \left[ \tfrac{1}{\left| B_t(0) \right| } \inf _{z \in {\mathbb {R}}^2} \# \big [ (\Lambda ' \times {\mathbb {Z}}) \cap B_t(z)\big ] \right] \liminf _{r \longrightarrow \infty } \frac{2}{\pi r^2} \int _s^r \left| B_t(0) \right| \,\frac{dt}{t} \\&=\inf _{t \ge s} \left[ \inf _{z \in {\mathbb {R}}^2} \tfrac{1}{\left| B_t(z) \right| } \# \big [ (\Lambda ' \times {\mathbb {Z}}) \cap B_t(z) \big ] \right] . \end{aligned}$$

Letting \(s \longrightarrow \infty \), we see that

$$\begin{aligned} D^+_{\mathrm {circ}}(\Lambda ) \ge \liminf _{t \longrightarrow \infty } \inf _{z \in {\mathbb {R}}^2} \tfrac{1}{\left| B_t(z) \right| } \# \big [ (\Lambda ' \times {\mathbb {Z}}) \cap B_t(z) \big ] = D^{-}(\Lambda ') = D^{-}(\Lambda ), \end{aligned}$$

where we used that the two-dimensional lower Beurling density of \(\Lambda ' \times {\mathbb {Z}}\) is \(D^{-}(\Lambda )\).

\(\square \)

We can now prove the uniqueness Theorem. The proof builds on [4, Lemma 5.1] and [5, Theorems 4.3 and 4.9].

FormalPara Proof of Theorem 2

We proceed by induction on m in (1).

In the case \(m=0\), the function g is a Gaussian \(g(x) = C'_0 e^{-a x^2}\), with \(a=\frac{\pi ^2}{\gamma }>0\). Let \(f=\sum _k c_k g(\cdot -k)\in V^\infty (g)\) with c bounded be non-zero, and denote by \(\Lambda \subseteq {\mathbb {R}}\) its zero set.

As shown in [5, Lemma 4.1], f possesses an extension to an entire function on \(\mathbb {C}\), satisfying the growth estimate \(|f(x+iy)| \lesssim e^{a y^2}\) for \(x,y \in {\mathbb {R}}\), where the implied constant depends on f. Let \(n \ge 0\) be the order of f at \(z=0\), and consider the analytic function

$$F(z) := C_1 z^{-n} f(z) e^{\frac{a}{2} z^2},$$

where \(C_1 \in \mathbb {C}\) is chosen so that \(F(0)=1\). Then F satisfies

$$\begin{aligned} \left| F(x+iy) \right| \le C e^{ay^2} e^{\frac{a}{2} (x^2-y^2)} = C e^{\frac{a}{2} (x^2+y^2)}, \quad x,y \in {\mathbb {R}}, \end{aligned}$$
(11)

for some constant \(C>0\). Moreover, the zero set of f is invariant under addition of \(i\tfrac{\pi }{a}{\mathbb {Z}}\) [5, Lemma 4.2]. Therefore, the set of complex zeros of F contains \(\big (\Lambda \setminus \{0\}\big ) + i\tfrac{\pi }{a}{\mathbb {Z}}\):

$$\begin{aligned} F \big (\lambda + i \tfrac{\pi }{a} k \big ) = 0, \quad \lambda \in \Lambda \setminus \{0\}, k \in {\mathbb {Z}}. \end{aligned}$$

By (6) with \(\alpha =\tfrac{\pi }{a}\), the zero-counting function of F,

$$\begin{aligned} n_F(t) := \# \{ z \in \mathbb {C}: F(z)=0, \left| z \right| \le t \}, \end{aligned}$$

satisfies

$$\begin{aligned} \limsup _{r \longrightarrow \infty } \frac{1}{r^2} \int _0^r \frac{n_F(t)}{t} dt \ge \frac{a}{2} D^+_{\mathrm {circ}}(\Lambda ). \end{aligned}$$
(12)

On the other hand, by Jensen’s formula combined with (11), for all \(r>0\),

$$\begin{aligned} \frac{1}{r^2} \int _0^r \frac{n_F(t)}{t} dt = \frac{1}{2\pi r^2} \int _0^{2\pi } \log |{F\big (r e^{i\theta }\big )}| d\theta \le \frac{\log (C)}{r^2} + \frac{a}{2}. \end{aligned}$$
(13)

By combining (12) and (13) we conclude that \(D^+_{\mathrm {circ}}(\Lambda ) \le 1\), as claimed. Note, in particular, that \(\Lambda \) has no accumulation points.

Inductive step. Let g be as in (1) and \(f=\sum _k c_k g(\cdot -k) \in V^\infty (g)\) non-zero. Without loss of generality we assume that, in (1), \(\delta _1, \ldots , \delta _m\) are non-zero. Since g is real-valued, by replacing \(c_k\) with \(\mathfrak {R}(c_k)\) or \(\mathfrak {I}(c_k)\) if necessary, we may assume that f is real-valued. Let \(g_1\) be given by (1), except that the product runs only up to \(m-1\). Then

$$\begin{aligned} f_1 := f + \delta _m f' = \sum _{k\in {\mathbb {Z}}} c_k g_1(\cdot -k) \end{aligned}$$

belongs to \(V^\infty (g_1)\) and is real-valued and non-zero, because the coefficients \(c_k\) are real and at least one of them is non-zero. We assume by inductive hypothesis that \(D^+_{\mathrm {circ}}(\{f_1=0\}) \le 1\).

Let \(\Lambda =\{f=0\}\) and \(\Gamma =\{f_1=0\}\). Suppose that \(\Lambda \) has no accumulation points. Assume for the moment that, in addition, f has a non-negative zero, and order the set of non-negative zeros of f increasingly:

$$\begin{aligned} \Lambda \cap [0,\infty )=\{\lambda _k : k=0,\ldots ,N\}, \end{aligned}$$

where \(N \in {\mathbb {N}}\cup \{0,\infty \}\), and we do not count multiplicities. Hence, \(\lambda _k >0\), if \(k>0\).

By Rolle’s Theorem applied to the differential operator \(I + \delta _m \partial _x\), there is a sequence of zeros of \(f_1\) that interlaces \(\Lambda \cap [0,\infty )\). (Indeed, we note that \(\frac{d}{dx} \big [ e^{x/\delta _m} f(x)\big ] = \frac{1}{\delta _m} e^{x/\delta _m} f_1(x)\) and apply the standard version of Rolle’s theorem; see [4, Lemma 5.1] or [5, Lemma 4.8] for details.) We parameterize that sequence as \(\Gamma ^+=\big \{\gamma _k : k=1,\ldots ,N\}\), with

$$\begin{aligned} \lambda _{k-1}< \gamma _k < \lambda _k, \quad k>0. \end{aligned}$$

The set \(\Gamma ^+\) is empty if \(N=0\). For \(t>0\), this ordering associates with each \(\lambda _k \in (0,t]\), \(k \not =0\), a distinct zero of \(f_1\), \(\gamma _k \in (0,t]\), such that the vertical segment through \(\gamma _k\) contained in the disk \(B_t(0)\) is longer than the corresponding segment through \(\lambda _k\):

$$\begin{aligned} \sum _{k>0, \lambda _k \in (0,t] } \sqrt{t^2 - \lambda _k^2}&\le \sum _{k >0, \gamma _k \in (0,t] } \sqrt{t^2 - \gamma _k^2} \le \sum _{\gamma \in \Gamma \cap (0,t] } \sqrt{t^2 - \gamma ^2}; \end{aligned}$$

see Fig. 1.

Fig. 1
figure 1

Each positive zero \(\lambda _k\) of f is assigned to a zero \(\gamma _k\) of \(f_1\) in such a way that the corresponding vertical segments contained in the disk \(B_t(0)\) become longer

We now add the term corresponding to \(\lambda _0\), and bound \(\sqrt{t^2 - \lambda _0^2} \le t\) to obtain

$$\begin{aligned} \sum _{\lambda \in \Lambda \cap [0,t] } \sqrt{t^2 - \lambda ^2} \le t + \sum _{\gamma \in \Gamma \cap (0,t] } \sqrt{t^2 - \gamma ^2}. \end{aligned}$$

The previous estimate is also trivially true is f has no non-negative zeros. Arguing similarly with the non-positive zeros of f we conclude that

$$\begin{aligned} \sum _{\lambda \in \Lambda \cap [-t,t] } \sqrt{t^2 - \lambda ^2} \le 2t + \sum _{\gamma \in \Gamma \cap [-t,t] } \sqrt{t^2 - \gamma ^2}. \end{aligned}$$

This shows that \(D^+_{\mathrm {circ}}(\{f=0\}) \le D^+_{\mathrm {circ}}(\{f_1=0\})\). The latter quantity is bounded by 1 by inductive hypothesis. Finally, if \(\Lambda =\{f=0\}\) has an accumulation point, Rolle’s theorem applied as before shows that so does \(\Gamma =\{f_1=0\}\), and, therefore, \(D^+_{\mathrm {circ}}(\{f_1=0\}) = \infty \) contradicting the inductive hypothesis. \(\square \)