ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P4.05 https://doi.org/10.26493/1855-3974.3114.d47 (Also available at http://amc-journal.eu) Complexity function of jammed configurations of Rydberg atoms* Tomislav Došlić Department of Mathematics, Faculty of Civil Engineering, University of Zagreb, Zagreb, Croatia and Faculty of Information Studies, Novo Mesto, Slovenia Mate Puljiz , Stjepan Šebek † , Josip Žubrinić Department of Applied Mathematics, Faculty of Electrical Engineering and Computing, University of Zagreb, Zagreb, Croatia Received 3 May 2023, accepted 4 March 2024, published online 25 September 2024 Abstract In this article, we determine the complexity function (configurational entropy) of jammed configurations of Rydberg atoms on a one-dimensional lattice. Our method con- sists of providing asymptotics for the number of jammed configurations determined by direct combinatorial reasoning. In this way we reduce the computation of complexity to solving a constrained optimization problem for the Shannon’s entropy function. We show that the complexity can be expressed explicitly in terms of the root of a certain polynomial of degree b, where b is the so-called blockade range of a Rydberg atom. Our results are put in a relation with the model of irreversible deposition of k-mers on a one-dimensional lattice. Keywords: Dynamic lattice systems, equilibrium lattice systems, complexity function, configurational entropy, jammed configuration, maximal packing, Rydberg atoms. Math. Subj. Class. (2020): 82B20, 82C20, 05B40, 05A15, 05A16 *It is a pleasure for us to thank Jean-Marc Luck and Pavel Krapivsky for fruitful exchanges during the concomi- tant elaboration of their preprint [43] and of the present work. T. Došlić gratefully acknowledges partial support by the Slovenian ARIS via Program P1-0383, grant no. J1-3002, and by COST Action CA21126 NanoSpace. Financial support of the Croatian Science Foundation (project IP-2022-10-2277) is gratefully acknowledged by S. Šebek. Lastly, we thank the anonymous referee for helpful comments that have led to improvements of the presentation of the article. †Corresponding author. E-mail addresses: tomislav.doslic@grad.unizg.hr (Tomislav Došlić), mate.puljiz@fer.unizg.hr (Mate Puljiz), stjepan.sebek@fer.unizg.hr (Stjepan Šebek), josip.zubrinic@fer.unizg.hr (Josip Žubrinić) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 2 Ars Math. Contemp. 24 (2024) #P4.05 1 Introduction Rydberg atom is a name given to an atom which has been excited into a high energy level so that one of its electrons is able to travel much farther from the nucleus than usual (up to 106 times more, see [22]). In physics community, Rydberg atoms have been intensely studied and have become a testing ground for various quantum mechanical problems in quantum information processing, quantum computation and quantum simulation [58]. See [29] for a comprehensive description of the physics of Rydberg atoms and their remarkable properties. Due to their large size, Rydberg atoms can exhibit very large electric dipole moments which results in strong interactions between two close Rydberg atoms. This causes a blockage effect that prohibits the excitation of an atom located close to an atom that is already excited to a Rydberg state [3, 12, 34, 37, 54, 63]. The simplest setting for studying Rydberg atoms and their blockage effect is on a finite one-dimensional lattice. In this setting, each atom occupies one site and each two excited atoms are at least b ≥ 1 sites apart. The positive integer b is referred to as the blockade range of the model. We will be interested in maximal (or jammed) configurations where no further atoms can be excited. Note that in such a configuration each two excited atoms are at most 2b sites apart. In physics literature, jammed states in similar deposition models have the interpretation of metastable states at low enough temperature and/or high enough density, and are referred to as valleys, pure states, quasi-states, and inherent structures [4, 5, 17, 33, 38, 50, 62]. The main question related to maximal configurations concerns the expected density of the atoms excited to a Rydberg state. There are two natural ways to interpret this question. One way to look at this problem is to consider the set of all the possible maximal configu- rations and to sample one such configuration at random (which implies that all the maximal configurations are equiprobable). This is referred to as the static (or equilibrium) model. The static model is usually described by the so-called complexity function (also known as configurational entropy), and the expected density of particles in a jammed configuration converges to the argument of the maximum of the complexity function. This is exactly the approach we take in this paper and our main result is the derivation of the mentioned complexity function. Static model can be compared with the random sequential adsorp- tion (RSA) model (also refered to as the dynamic model) where initially all atoms are in the ground state, and are excited sequentially, at random, until a jammed configuration is reached. The expected density of excited atoms with this underlying probability space is called jamming limit. Assumption that the two models result in the same distribution of maximal configurations has come to be known as Edwards’s flatness hypothesis (see [2] for a recent review). However, there seems to be no a priori reason for the two models to have similar properties. It is interesting to note that there is a long history to the question of how similar the two models are (see e.g. the discussion in [8, page 681], or, in a more subtle continuum context, how a similar confusion of different probability models led to some extended discussion over a false “proof” [53]). The dynamic version of the problem was already studied in literature. In [28, 42, 48] it was found that the jamming limit is ρb-Ryd∞ = ∫ 1 0 exp −2 b∑ j=1 1− yj j  dy. The jamming limit was also computed for an equivalent model of deposition of linear T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 3 polymers (k-mers) in [44, §7.1] ρk-mer∞ = k ∫ ∞ 0 exp −u− 2 k−1∑ j=1 1− e−ju j  du = k ∫ 1 0 exp −2 k−1∑ j=1 1− yj j  dy. The equivalence of models is reflected in the fact that ρk-mer∞ = k · ρb-Ryd∞ for b = k − 1. In the static model, it all comes down to counting the maximal configurations. It is known that in similar models, the number of different maximal configurations with pre- scribed density 0 ≤ ρ ≤ 1 tends to grow exponentially with the length L of configuration, see [9, 10, 14, 13, 15, 16, 18, 23, 27, 35, 40, 45, 46, 49, 52, 55, 56, 59]. Denoting this number by JL(ρ), it is common to describe it using the so-called complexity function (also called configurational entropy) f(ρ) for which it holds that JL(ρ) ∼ eLf(ρ). It turns out (see e.g. Figure 9) that the density ρb-Ryd? maximizing the complexity function is slightly different than the expected density (jamming limit) of the dynamic model. This falsifies the above mentioned Edward’s flatness hypothesis. Recall that ρb-Ryd? is the limit (as L tends to infinity) of the most probable densities in the equilibrium models that assign equal probabilities to all jammed configurations. Our main goal is to compute the complexity f(ρ) of jammed configurations of Rydberg atoms using direct combinatorial reasoning. The problem reduces to solving a constrained optimization problem for the Shannon’s entropy function. We show that the complexity function can be expressed explicitly in terms of the root of a certain polynomial of degree b. This work has been carried out simultaneously with [43]. The authors there introduce a novel method for determining the same complexity function. Their method is inspired by the theory of renewal processes. The described model of Rydberg atoms on a one-dimensional lattice is equivalent to a number of other models already present in the literature. The case b = 1 is the famous model introduced by Flory [26] describing the mechanism of vinyl polymerization. This is in turn essentially equivalent to the Page-Rényi car parking problem [31, 51] (which is a discrete version of the famous model introduced by Rényi in [57]) describing the jammed configurations of cars of length 2. The equivalence is obtained by replacing each excited atom with a car taking up both the atom’s and its right neighbor’s site. Clearly, this only works for configurations not having an excited atom at the rightmost site. This means that the total number of jammed configurations is actually different in these two models, but only up to a constant factor, which does not affect the shape of the complexity function of these models. In chemistry, this model appears in the context of the irreversible deposition of dimers [26, 36], and in graph theory, the jammed configurations correspond to maximal matchings in a path graph, see [21]. Similarly, the general case b > 1 corresponds to irreversible deposition of k-mers (k = b + 1) in a linear polymer of length L. The equivalence (again, up to a constant factor) is obtained by replacing each excited atom with a polymer taking up b + 1 consecutive sites, starting from the atom’s position, see Figure 1. In this, and all the following figures, bullets (•) represent Rydberg atoms (in the Rydberg model) or occupied sites (in the k-mer deposition model), while empty bullets (◦) represent neutral atoms (in the Rydberg model) or vacant sites (in the k-mer deposition model). Notice that the gaps between adjacent k-mers in jammed configurations of this deposition model are of size at most k − 1. This equivalence allows us to easily transfer our results on Rydberg atoms to the setting of k- 4 Ars Math. Contemp. 24 (2024) #P4.05 mer deposition model. The problem of irreversible deposition of k-mers was extensively studied in the literature, see [1, 6, 24, 32, 42, 44, 61]. ◦ • ◦ ◦ ◦ ◦ • ◦ ◦ ◦ ◦ ◦ ◦ • ◦ ◦ ◦ • ◦ ◦ ◦ • ◦ ◦ ◦ ◦ ◦ ◦ • ◦ ◦ ◦ • ◦ ◦ ◦ ◦ ◦ • ◦ ◦ ◦ ◦ • • • • ◦ • • • • ◦ ◦ ◦ • • • • • • • • • • • • ◦ ◦ ◦ • • • • • • • • ◦ ◦ • • • • Figure 1: A jammed configuration of Rydberg atom model with blockade range b and the corresponding jammed configuration of the k-mer deposition model when b = 3, k = 4. In graph theory, the k-mer deposition model is equivalent to Pk-packings of a path graph PL, and jammed configurations in the former correspond to maximal packings in the latter. The maximal Pk-packings of PL were previously studied in [20]. Another equivalent formulation of the Rydberg atom model appeared recently in [19, §3.2.1] where the authors of the present paper considered the settlement model consisting of k-story buildings on a one-dimensional tract of land. The tract of land is oriented east- west and each story of each building has to receive the sunlight from both east and west. The rest of the paper is organized as follows. In Section 2 we calculate the asymp- totics for the number of jammed configurations in the model of Rydberg atoms, which is expressed in terms of the maximum of the Shannon’s entropy function over a certain finite set determined by the constraints of the model. In Section 3 we use these results in order to obtain the formula for the associated complexity function. We derive the expression for the complexity f(ρ) which, for a chosen density ρ, depends explicitly on a positive root of a certain polynomial whose degree coincides with the blockade range of the model. Further on, in Section 4, we put our findings in relation with the model for the deposition of k-mers on the linear polymer and draw conclusions from the obtained results. There, we also provide some results on the qualitative properties of the maximizers of mentioned complexity functions, for various blockade ranges b, and put them in comparison with their counterparts in the theory of RSA. Finally, in Section 5 we recapitulate our findings and indicate several possible directions of future research. Notation We write ML ∼ NL if the two positive sequences (ML)L and (NL)L have the same growth, as L→∞, up to a sub-exponential factor, i.e. if lim L→∞ lnML − lnNL L = 0. 2 Jammed configurations of Rydberg atoms As already stated in the introduction, the main goal of this paper is to compute the complex- ity function f(ρ) of jammed configurations of Rydberg atoms. Crucial step towards obtain- ing a complexity function of such models in general is to inspect the set of all jammed con- figurations of a model. Each configuration is a binary 0/1 sequence which we sometimes interpret as a sequence of empty/occupied sites or, in Rydberg model, as neutral/excited atoms. The total number of all configurations of length L in the model is denoted by JL. The total number of configurations of lengthL consisting ofN ones (occupied sites, excited T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 5 atoms) is denoted by JN,L. The density (saturation, coverage) of any such configuration of length L with N ones is defined as N/L ∈ [0, 1]. In order to determine the complexity function, it is not enough to work only with JL. We need to be more precise. We need to know the behavior of the number of different jammed configurations of length L, where precisely N atoms are excited to the Rydberg state. The main result of this section (see Lemma 2.4) provides asymptotics of the quantity JN,L for Rydberg atom model. Let us first consider several concrete examples of jammed configurations of our model to get a better feeling of their possible shapes. Figure 2 displays three different jammed configurations in the chain of L = 16 atoms, where the blockade range b is equal to two, i.e. each two excited atoms are at least two sites apart. Since Rydberg atoms in a jammed • ◦ ◦ • ◦ ◦ • ◦ ◦ • ◦ ◦ • ◦ ◦ • ◦ • ◦ ◦ • ◦ ◦ ◦ ◦ • ◦ ◦ • ◦ ◦ • • ◦ ◦ ◦ ◦ • ◦ ◦ ◦ ◦ • ◦ ◦ ◦ • ◦ Figure 2: Three jammed configurations in the chain of L = 16 atoms with blockade range b = 2. The number of Rydberg atoms in these configurations is N = 6, 5, 4 (from top to bottom). configuration are separated by clusters of empty sites whose length is at least b (so that the constraint imposed by the blockage effect is satisfied), and at most 2b (since we can excite another atom in the middle of an empty range of size 2b + 1, hence such a configuration would not be jammed), it is easy to see that it holds⌈ L 2b+ 1 ⌉ ≤ N ≤ ⌈ L b+ 1 ⌉ , (2.1) where N is the number of excited atoms, L is the length of the configuration, and b is the blockade range. In the particular case of L = 16 and b = 2, this implies that 4 ≤ N ≤ 6. Hence, Figure 2 shows one jammed configuration for each possible value ofN . Notice that relation (2.1) implies that 1 2b+ 1 − 1 L < N L ≤ 1 b+ 1 , (2.2) and this in turn implies that in the limit, as L → ∞, the density ρ = N/L, of Rydberg atoms in jammed configurations, lies within the bounds 1 2b+ 1 ≤ ρ ≤ 1 b+ 1 . (2.3) As a first result in the direction of better understanding the double sequence JN,L for Rydberg atom model, we provide the bivariate generating function for this sequence in the general case of blockade range b ≥ 1. Lemma 2.1. The bivariate generating function of the sequence JN,L associated with jammed configurations of Rydberg atoms, when the blockade range is equal to b, is given by Fb(x, y) = ∑ L ∑ N JN,Lx NyL = (1− y)2 + xy − xyb+1 − xyb+2 + xy2b+2 (1− y)(1− y − xyb+1 + xy2b+2) . 6 Ars Math. Contemp. 24 (2024) #P4.05 Proof. As already mentioned, configurations of Rydberg atoms can be represented as 0/1 sequences. Due to the fact that we can determine whether the blockage effect has been taken into account, and whether the configuration represented with such a sequence is jammed, just by inspecting finite size patches of a given sequence, we can apply the so- called transfer matrix method (see [60, §4.7] or [25, §V], and also [47, §2–4]). This is a well known method for counting words of a regular language. Since Rydberg atoms in a jammed configuration are separated with at least b, and at most 2b neutral atoms, every jammed configuration will be composed of blocks that start with a Rydberg atom and then have a cluster of neutral atoms of length between b and 2b. Such blocks are displayed in Figure 3. • ◦ · · · ◦︸ ︷︷ ︸ b atoms • ◦ · · · ◦︸ ︷︷ ︸ b+1 atoms · · · • ◦ · · · ◦︸ ︷︷ ︸ 2b atoms Figure 3: Building blocks of jammed configurations of Rydberg atoms with blockade range b. These building blocks are encoded with the polynomial pb(x, y) = xy b+1 + xyb+2 + · · ·+ xy2b+1. Now we only need to take care of the beginning and the end of jammed configurations. Notice that in front of the first block we can have some neutral atoms. More precisely, the number of neutral atoms that can appear at the left end of the jammed configuration is between 0 and b. These starting blocks are encoded with the polynomial sb(x, y) = 1 + y + y 2 + · · ·+ yb. Similarly, after the last block from the set of blocks shown in Figure 3 (if there are any, i.e. if we want to have more than just one atom in the Rydberg state), we need to have a block that again starts with a Rydberg atom, and then has a cluster of neutral atoms of length between 0 and b. These ending blocks are encoded with the polynomial eb(x, y) = xy + xy 2 + · · ·+ xyb+1. Notice that each of the blocks shown in Figure 3 can be glued to any other block listed in this figure. This implies that we do not even need to work with powers of the transfer matrix, but we can directly take powers of the polynomial pb(x, y) in order to obtain the desired bivariate generating function. A simple calculation gives Fb(x, y) = 1 + ∞∑ n=0 sb(x, y) · pb(x, y)n · eb(x, y) = 1 + sb(x, y) · eb(x, y) 1− pb(x, y) = (1− y)2 + xy − xyb+1 − xyb+2 + xy2b+2 (1− y)(1− y − xyb+1 + xy2b+2) . T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 7 Remark 2.2. By using the same technique, we can easily compute the bivariate generating function enumerating the number of jammed configurations of prescribed length, and with some fixed number of occupied sites, in the k-mer deposition model. The building blocks here are composed of a cluster of k consecutive sites occupied by a single k-mer, followed by a cluster of empty sites of length between 0 and k − 1 (see Figure 4). These building • · · · •︸ ︷︷ ︸ k sites • · · · •︸ ︷︷ ︸ k sites ◦ • · · · •︸ ︷︷ ︸ k sites ◦ ◦ · · · • · · · •︸ ︷︷ ︸ k sites ◦ · · · ◦︸ ︷︷ ︸ k−1 sites Figure 4: Building blocks of jammed configurations of k-mer deposition model. blocks are encoded with a polynomial pk(x, y) = x kyk + xkyk+1 + · · ·+ xky2k−1, where x is again a formal variable associated with the number of occupied sites, and y is a formal variable associated with the length of a configuration. Similarly as in the case of the Rydberg atom model, at the left end of a jammed configuration, we can have a cluster of vacant sites of length between 0 and k − 1. These starting blocks are encoded with the polynomial sk(x, y) = 1 + y + y 2 + · · ·+ yk−1. It is clear that we can end a jammed configuration with any of the building blocks shown in Figure 4, so we can set ek(x, y) = 1. Using again the fact that each of the blocks from Figure 4 can be glued to any other block listed in that figure, we can work directly with powers of the polynomial pk(x, y) to obtain Fk(x, y) = ∞∑ n=0 ak(x, y) · pk(x, y)n = ak(x, y) 1− pk(x, y) = 1− yk 1− y − xkyk + xky2k . (2.4) Notice that we are not adding 1 to the bivariate generating function in (2.4). The reason is that starting with a cluster of 0 vacant sites and setting n = 0 already counts the empty configuration. The sequence JN,L has already been studied in the literature, but in the context of maximal Pk-packings of a path graph PL (see [20]). The bivariate generating function enumerating the total number of maximal k-packings in PL, with exactly N copies of Pk, is given in [20, Corollary 2.4], and the only difference between that bivariate generating function and the one we obtained in (2.4), is that x is not raised to power k. The reason is that the author in [20] is interested in the number of copies of Pk (i.e. the number of de- posited k-mers) in jammed configurations, and we are interested in the total number of sites occupied by those deposited k-mers. The bivariate generating function from (2.4) is also obtained in [43, formula (5.3)], where authors use a novel approach inspired by the theory of renewal processes. Using the same technique, they also obtain the bivariate generating function which coincides with the one we obtained in Lemma 2.1, which enumerates the total number of jammed configurations of length L of Rydberg atoms with blockade range b, with precisely N excited atoms (see [43, formula (6.5)]). It is easy to see from the bivariate generating function from Lemma 2.1 that, for b = 2, J16 = 96 (i.e. there are 96 jammed configurations in the chain of L = 16 atoms, when 8 Ars Math. Contemp. 24 (2024) #P4.05 the blockade range is b = 2). Out of those 96 jammed configurations, 45 of them have 4 Rydberg atoms (J4,16 = 45), 50 of them have 5 Rydberg atoms (J5,16 = 50), and only one has 6 Rydberg atoms (J6,16 = 1). This particular one is exactly the first jammed configuration shown in Figure 2. We could now proceed like the authors in [43] and use the bivariate generating function developed in Lemma 2.1 to obtain the complexity function of jammed configurations of Rydberg atoms by means of the Legendre transform. However, we will use a direct combi- natorial argument. To this end, we introduce a slightly different way of counting jammed configurations in the Rydberg model with blockade range b, than the one introduced in Lemma 2.1. Denote with B the block of b + 1 adjacent atoms where only the first one is excited to the Rydberg state (see Figure 5). Using again the fact that each two Rydberg B = • ◦ ◦ · · · ◦︸ ︷︷ ︸ b atoms Figure 5: Block consisting of b+1 adjacent atoms where only the first one is excited to the Rydberg state. atoms have at least b and at most 2b neutral atoms separating them, it is clear that every jammed configuration consists of blocks B separated by clusters of neutral atoms of length 0 ≤ a ≤ b (see Figure 6). Denote by Ma the number of gaps with a neutral atoms. The ◦ · · · ◦︸ ︷︷ ︸ a1 atoms B ◦ · · · ◦︸ ︷︷ ︸ a2 atoms B ◦ · · · ◦︸ ︷︷ ︸ a3 atoms B · · ·B ◦ · · · ◦︸ ︷︷ ︸ aN atoms B Figure 6: The shape of jammed configurations in Rydberg model with blockade range b and exactly N Rydberg atoms, ending with a block B (displayed in Figure 5). Gaps between blocks B, and in front of the first block B, consist of neutral atoms and are of length 0 ≤ ai ≤ b. total number of jammed configurations of the shape shown in Figure 6, with L atoms in total, out of which precisely N atoms are excited to the Rydberg state, is given as( N M0,M1, . . . ,Mb ) = N !∏ 0≤a≤bMa! , (2.5) with Ma satisfying b∑ a=0 Ma = N, (2.6) b∑ a=0 aMa = L− (b+ 1)N. (2.7) The constraint (2.6) expresses that the total number of gaps is N . Notice that we have N blocks B (since we want to have precisely N Rydberg atoms), and that gaps of size 0 ≤ a ≤ b can be added in front of the first block B, and between each two blocks B. T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 9 The constraint (2.7) implies that the total number of neutral atoms is L − N . Clearly we need L − N neutral atoms in addition to N Rydberg atoms to have a configuration of length L. Equation (2.5) accounts for the jammed configurations ending precisely on B. There are also jammed configurations where the last block B is truncated, and there are only 0 ≤ c < b neutral atoms after the last atom excited to the Rydberg state. The contribution of such jammed configurations to the value of JN,L is comparable to (2.5), but since complexity function ignores sub-exponential factors, it suffices to determine the asymptotics of the sum JN,L ∼ ∑ (M0,M1,...,Mb)∈RN,L ( N M0,M1, . . . ,Mb ) , (2.8) where RN,L = {(M0,M1, . . . ,Mb) ∈ Nb+10 :M0 +M1 + · · ·+Mb = N and M1 + 2M2 + · · ·+ bMb = L− (b+ 1)N}. (2.9) We write H for the Shannon’s entropy function given as H(p0, p1, . . . , pb) = − b∑ i=0 pi ln pi, (2.10) where pi ≥ 0, for 0 ≤ i ≤ b, and p0 + p1 + · · ·+ pb = 1. Remark 2.3. In case pi = 0 for some i, we set 0 · ln 0 = 0. The following lemma is the key result of this section, and it constitutes a crucial step in computing the complexity function of our model as it provides the asymptotics of JN,L in terms of the maximum of the entropy function. Lemma 2.4. JN,L ∼ exp ( L · max (M0,M1,...,Mb)∈RN,L N L ·H ( M0 N , M1 N , . . . , Mb N )) , as L→∞. where the set RN,L is defined in (2.9), and the function H is defined in (2.10). Proof. Note that max (M0,M1,...,Mb)∈RN,L ( N M0,M1, . . . ,Mb ) ≤ ∑ (M0,M1,...,Mb)∈RN,L ( N M0,M1, . . . ,Mb ) ≤ |RN,L| max (M0,M1,...,Mb)∈RN,L ( N M0,M1, . . . ,Mb ) . As the number |RN,L| of terms in the sum is at most (N + 1)b+1 ≤ (L+ 1)b+1, which is polynomial in L, the sum, asymptotically, grows as its largest term. It is, therefore, enough to determine the asymptotics of JN,L ∼ max (M0,M1,...,Mb)∈RN,L ( N M0,M1, . . . ,Mb ) , as L (and N)→∞. 10 Ars Math. Contemp. 24 (2024) #P4.05 By following the proof of Lemma 2.2 in [11] we can conclude that( N + b b )−1 NN M0 M0M1 M1 · · ·MbMb ≤ ( N M0,M1, . . . ,Mb ) ≤ N N M0 M0M1 M1 · · ·MbMb . Note that in case any Ma is zero, the expression 00 is to be interpreted as 1. Since ( N+b b ) is of polynomial growth, we get( N M0,M1, . . . ,Mb ) ∼ N N M0 M0M1 M1 · · ·MbMb = ( N M0 )M0 ( N M1 )M1 · · · ( N Mb )Mb , (2.11) as N →∞. Note that( N M0 )M0 ( N M1 )M1 · · · ( N Mb )Mb = exp ( N ·H ( M0 N , M1 N , . . . , Mb N )) . Hence JN,L ∼ max (M0,M1,...,Mb)∈RN,L exp ( N ·H ( M0 N , M1 N , . . . , Mb N )) , as L→∞, and consequentially JN,L ∼ exp ( L · max (M0,M1,...,Mb)∈RN,L N L ·H ( M0 N , M1 N , . . . , Mb N )) , as L→∞, which is exactly what we wanted to prove. Remark 2.5. One could obtain the asymptotics in (2.11) from Stirling’s approximation N ! ∼ (N/e)N , as N →∞, where sub-exponential factors are ignored. 3 Complexity function of jammed configurations of Rydberg atoms In this section we compute the complexity function, sometimes referred to as configura- tional entropy, of jammed configurations of Rydberg atoms. We first recall the definition of complexity function of a certain model. Definition 3.1. For a fixed density ρ ∈ [0, 1], let JbρLc,L denote the number of configu- rations of length L with density bρLc /L ≈ ρ. The complexity function f : [0, 1] → R is then defined as f(ρ) = lim L→∞ ln JbρLc,L L , (3.1) for each ρ ∈ [0, 1] for which this limit exists. Remark 3.2. If the limit above does not exist for a certain ρ, one can still define (upper) complexity at that point by replacing lim in the definition with lim sup. And if there are no configurations with a certain density ρ, we still write f(ρ) = 0. Remark 3.3. This definition implies that the number of configurations with the density bρLc /L ≈ ρ grows as eLf(ρ) for large L. T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 11 The guiding idea behind introducing the complexity function is to describe what portion of the total number of configurations take up configurations with a particular density. The problem is that, as L grows to infinity, the actual proportions tend to the delta distribution concentrated on the ‘most probable’ density ρ?. As an example, the distribution of densities (the sum of digits divided by the length) of binary sequences of length L is a symmetric binomial distribution re-scaled to the interval [0, 1]. The limiting distribution is then the delta distribution δ0.5 which is, essentially, the consequence of the law of large numbers. This convergence to a delta distribution results from the fact that the number of config- urations with a certain density grows exponentially with a rate that depends on the density. For large L, the number of configurations with density having the largest rate overtakes, in proportion, configurations having any other density. The complexity function then quan- tifies the distribution of all configurations with respect to their densities in a more refined way. Another consequence of the fact that the number of configurations having density with the largest rate dominates, in proportion, any other density is that the total number of all configurations grows at the same exponential rate as the number of configurations having this ‘most probable density’. To be precise, if ρ? denotes the density at which the com- plexity function f attains its maximum and if JL is the total number of all configurations of length L, then JL ∼ eLf(ρ?) for large L. Remark 3.4. In Lemma 2.1 we derived the generating function for the sequence JN,L within the Rydberg atom model. Plugging x = 1 into this generating function gives the generating function for JL, the total number of configurations of length L in Rydberg atom model Fb(1, y) = (1− y)2 + y − yb+1 − yb+2 + y2b+2 (1− y)(1− y − yb+1 + y2b+2) = 1 + y(1 + y + · · ·+ yb)(1 + y + · · ·+ yb−1) 1− yb+1(1 + y + · · ·+ yb) . From here, we can infer the asymptotics of JL for large L by inspecting the roots of the polynomial 1− yb+1(1+ y+ · · ·+ yb) in the denominator. More precisely, if yb is the root with the smallest modulus, then the logarithm of wb = |yb|−1 gives the exponential growth rate of the sequence JL JL ∼ wLb = eL lnwb . The discussion in the previous paragraph now implies the relation f(ρb-Ryd? ) = lnwb. The following theorem is the main result of this paper and provides an elegant expres- sion for the complexity function of jammed configurations of Rydberg atoms f(ρ) in terms of a root of a certain polynomial. Theorem 3.5. The complexity function of jammed configurations of Rydberg atoms with blockade range b ∈ N is given as f(ρ) = { ρ [ − ln 1−z 1−zb+1 − ( 1 ρ − (b+ 1) ) ln z ] , if 12b+1 < ρ ≤ 1b+1 , 0, otherwise, 12 Ars Math. Contemp. 24 (2024) #P4.05 where z ≥ 0 is a real root of the polynomial p(z) = b∑ i=0 ( i+ b+ 1− 1 ρ ) zi (3.2) for which the expression f(ρ) is the largest. Remark 3.6. When 12b+1 < ρ < 1 b+1 the leading coefficient of the polynomial p(z) given in (3.2) is positive, while the constant term is negative. This guaranties the existence of at least one positive real root z > 0. If ρ = 1b+1 , then z = 0 is the root of p(z) and the formula gives f( 1b+1 ) = 0. Remark 3.7. Since (3.2) is a polynomial of degree b, it is possible to find its roots explicitly for b ≤ 4 and numerically for b > 4. The explicit expression for the complexity in case b = 1 is f1-Ryd(ρ) = ρ ln ρ− (1− 2ρ) ln(1− 2ρ)− (3ρ− 1) ln(3ρ− 1), and for b = 2 f2-Ryd(ρ) = (3ρ− 1) ln √ −44ρ2 + 24ρ− 3− 4ρ+ 1 10ρ− 2 − ρ ln −350ρ3 + (25ρ2 − 10ρ+ 1) √ −44ρ2 + 24ρ− 3 + 215ρ2 − 44ρ+ 3 ρ2 √ −44ρ2 + 24ρ− 3− 134ρ3 + 57ρ2 − 6ρ . In the case b = 1, the function f1-Ryd(ρ) recovers the result from [44, formula (7.20)] and [41, §VII]. The graphs of the complexity function of jammed configurations of Rydberg atoms with blockade range 1 ≤ b ≤ 10 are given in Figure 7. In that figure we also see that, for each b, the maximum of the complexity function matches lnwb, the growth rate of all jammed configurations. This was already discussed in Remark 3.4. Proof of Theorem 3.5. Recall that in (2.2) we showed that 1 2b+ 1 − 1 L < N L ≤ 1 b+ 1 , and therefore, there are no jammed configurations with densities ρ > 1b+1 nor with densities ρ < 12b+1 , for sufficiently large L. Thus, f(ρ) = 0 when ρ > 1 b+1 or ρ < 1 2b+1 . In case ρ = 12b+1 , it is not hard to see that the number of configurations Jb L2b+1c,L is Jb L2b+1c,L = { 1, if (2b+ 1) | L, 0, otherwise. This implies f( 12b+1 ) = 0 by the definition of complexity. In the remainder, we fix 12b+1 < ρ ≤ 1b+1 . By Lemma 2.4, and by using the definition of the complexity function (3.1), we have f(ρ) = lim L→∞ max (M0,M1,...,Mb)∈RN,L N L ·H ( M0 N , M1 N , . . . , Mb N ) , (3.3) T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 13 1 2 1 3 1 4 1 5 1 6 1 7 1 8 1 9 0 ρ 0 0.05 0.1 0.3 lnw1 = lnw2 lnw3 lnw4 lnw5 lnw6 lnw7 lnw8 lnw9 lnw10 f (ρ ) b = 1 b = 2 b = 3 b = 4 b = 5 b = 6 b = 7 b = 8 b = 9 b = 10 Figure 7: The complexity function of jammed configurations of Rydberg atoms with block- ade range 1 ≤ b ≤ 10. where N = bρLc, provided that this limit exists. By rewriting (M0,M1, . . . ,Mb) ∈ RN,L as M0 N ≥ 0, M1 N ≥ 0, . . . , Mb N ≥ 0 M0 N + M1 N + · · ·+ Mb N = 1 M1 N + 2 M2 N + · · ·+ bMb N = L N − (b+ 1) and denoting pi = MiN ∈ 1bρLcZ, the complexity (3.3) can be written as f(ρ) = lim L→∞ max (p0,p1,...,pb)∈ 1bρLcRbρLc,L ρ̂H (p0, p1, . . . , pb) , (3.4) where ρ̂ = ρ̂(L) = NL = bρLc L . We claim that this limit exists and is equal to the maximum of the constrained optimization problem max p0,p1,...,pb≥0 p0+p1+···+pb=1 p1+2p2+···+bpb= 1ρ−(b+1) ρH (p0, p1, . . . , pb) , (3.5) where pi ∈ R are no longer required to be fractions. 14 Ars Math. Contemp. 24 (2024) #P4.05 We argue as follows. Denote by (p∗0, p ∗ 1, . . . , p ∗ b) the point at which the maximum in (3.5) is attained. For each L ∈ N, let (p0(L), p1(L), . . . , pb(L)) be the point at which maximum in (3.4) is attained. Clearly, ρ̂H (p0(L), p1(L), . . . , pb(L)) ≤ ρ̂H(p∗0, p∗1, . . . , p∗b) ≤ ρH(p∗0, p∗1, . . . , p∗b). The first inequality follows by substituting ρ̂ for ρ in (3.5) and the fact that one is now optimizing over a larger set. The second inequality follows from ρ̂ ≤ ρ. Note that the right hand side no longer depends on L, and thus lim sup L→∞ ρ̂H (p0(L), p1(L), . . . , pb(L)) ≤ ρH(p∗0, p∗1, . . . , p∗b). Next, for each L ∈ N, we consider the point (t0(L), t1(L), . . . , tb(L)) ∈ 1bρLcRbρLc,L, which is closest to the to the optimizer (p∗0, p ∗ 1, . . . , p ∗ b). Note that, due to the density argument, (t0(L), t1(L), . . . , tb(L)) → (p∗0, p∗1, . . . , p∗b) as L → ∞. This, along with the continuity of H and the fact that ρ̂→ ρ implies the lower bound ρH(p∗0, p ∗ 1, . . . , p ∗ b) = lim L→∞ ρ̂H(t0(L), t1(L), . . . , tb(L)) ≤ ≤ lim inf L→∞ ρ̂H (p0(L), p1(L), . . . , pb(L)) . Putting everything together completes the argument that the limit f(ρ) = lim L→∞ ρ̂H (p0(L), p1(L), . . . , pb(L)) exists and that the complexity function is f(ρ) = ρH(p∗0, p ∗ 1, . . . , p ∗ b) = max p0,p1,...,pb≥0 p0+p1+···+pb=1 p1+2p2+···+bpb= 1ρ−(b+1) ρ ·H (p0, p1, . . . , pb) . In order to obtain the expression for complexity f(ρ), it only remains to solve the con- strained optimization problem (3.5). We define the Lagrangian function L(p0, . . . , pb;λ, µ) = ρ ·H(p0, p1, . . . , pb)− λ(p0 + p1 + · · ·+ pb − 1) − µ(p1 + 2p2 · · ·+ bpb − 1 ρ + (b+ 1)), and find the stationary point by solving the system −ρ(ln pi + 1)− λ− µi = 0, for i = 0, 1, . . . , b; p0 + p1 + · · ·+ pb = 1; p1 + 2p2 · · ·+ bpb = 1 ρ − (b+ 1). (3.6) By multiplying i-th of the first (b+ 1) equations by pi and adding them together we get −ρ b∑ i=0 (pi ln pi + pi)− λ b∑ i=0 pi − µ b∑ i=0 ipi = 0, T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 15 and from here we obtain the expression for complexity in terms of the Lagrange multipliers λ and µ which solve the system (3.6) f(ρ) = ρH(p0, p1, . . . , pb) = ρ+ λ+ µ ( 1 ρ − (b+ 1) ) . (3.7) Subtracting successive equations in (3.6) we get −ρ(ln pi − ln pi−1)− µ = 0, or equivalently pi pi−1 = e−µ/ρ. Therefore pi = p0e−µi/ρ, for i = 1, . . . , b. From the very first equation in (3.6) we get p0 = e −λ/ρ−1, and the whole system (3.6) now reduces to just two equations e−λ/ρ−1 b∑ i=0 e−µi/ρ = 1; (3.8) e−λ/ρ−1 b∑ i=0 ie−µi/ρ = 1 ρ − (b+ 1). (3.9) Setting z = e−µ/ρ, and eliminating e−λ/ρ−1 from equations (3.8) and (3.9), gives a single polynomial equation of degree b bzb + (b− 1)zb−1 + · · ·+ 2z2 + z = [ 1 ρ − (b+ 1) ] (zb + zb−1 + · · ·+ z + 1), (3.10) which can be written as p(z) = 0 where p(z) is given in (3.2). Now, in order to obtain the complexity, all we need is, for a fixed 12b+1 < ρ < 1 b+1 , to find a real root z > 0 of the polynomial p(z) for which the expression (3.7) is the largest. The case ρ = 1b+1 , which gives z = 0, has to be treated separately. From relation z = e−µ/ρ and equation (3.8) we have µ = −ρ ln z; λ = −ρ ( 1 + ln 1− z 1− zb+1 ) . (3.11) Plugging (3.11) into (3.7), gives the complexity expressed in terms of the root of p(z) f(ρ) = ρ [ − ln 1− z 1− zb+1 − ( 1 ρ − (b+ 1) ) ln z ] . Lastly, in case ρ = 1b+1 , already from the last two equations in (3.6) we can con- clude p1 = p2 = · · · = pb = 0 and p0 = 1. This immediately gives f(ρ) = 0 as H(1, 0, 0, . . . , 0) = 0, completing the proof. 16 Ars Math. Contemp. 24 (2024) #P4.05 Remark 3.8. Using the standard summation formulas, we can rewrite (3.10) as bzb+2 − (b+ 1)zb+1 + z (1− z)2 = [ 1 ρ − (b+ 1) ] 1− zb+1 1− z , (3.12) or equivalently[ (2b+ 1)− 1 ρ ] zb+2 − [ (2b+ 2)− 1 ρ ] zb+1 − [ b− 1 ρ ] z + [ (b+ 1)− 1 ρ ] = 0. As discussed in the introduction, the complexity function is associated to equilibrium (or static) models of a certain phenomena and ρ?, the point at which the complexity func- tion attains its maximum, is interpreted as the expected and most probable density observed in such a model. This value ρ? is sometimes called the equilibrium density of the model and Theorem 3.9 below shows how to calculate it. A different (and perhaps more natural) way to look at Rydberg atom model is dynamically, within the framework of random sequential adsorption (RSA). Initially neutral atoms are sequentially and at random excited (obeying the blockade range constraint) until the jammed configuration is reached. The expected density of the reached jammed configuration (the jamming limit) in this dynamical version of the model, denoted by ρb-Ryd∞ , was computed in [42, §IV] ρb-Ryd∞ = ∫ 1 0 exp −2 b∑ j=1 1− yj j  dy. It is interesting to compare ρb-Ryd? and ρ b-Ryd ∞ for different blockade ranges b. Even though 0 20 40 60 80 100 b 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45 ρb-Ryd∞ ρb-Ryd? Figure 8: Comparison of ρb-Ryd? and ρ b-Ryd ∞ for 1 ≤ b ≤ 99. they are not the same, they seem to match quite nicely, see Figure 8. Additionally, as one would expect, they both tend to zero for large b. One can see their differences more clearly in Figure 9. This violation of Edwards flatness hypothesis is even more pronounced when one inspects the asymptotics of the two sequences more closely. In Figure 10 we see the graph of quantities b · ρb-Ryd? and b · ρb-Ryd∞ . T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 17 1 2 1 3 ρ 1-Ryd ? ρ1-Ryd∞ ρ 0.00 0.05 0.10 0.15 0.20 0.25 f1-Ryd(ρ) 1 6 1 11 ρ 5-Ryd ? ρ5-Ryd∞ ρ 0.00 0.05 0.10 0.15 0.20 f5-Ryd(ρ) 1 21 1 41 ρ 20-Ryd ? ρ20-Ryd∞ ρ 0.00 0.02 0.04 0.06 0.08 0.10 f20-Ryd(ρ) 1 51 1 101 ρ 50-Ryd ? ρ50-Ryd∞ ρ 0.00 0.01 0.02 0.03 0.04 0.05 f50-Ryd(ρ) Figure 9: Complexity function of Rydberg atom model with blockade range b, for b ∈ {1, 5, 20, 50}. Also plotted in each graph are the equilibrium density ρb-Ryd? and the jam- ming density ρb-Ryd∞ . 0 20 40 60 80 100 b 0.40 0.45 0.50 0.55 0.60 0.65 0.70 0.75 0.80 b · ρb-Ryd∞ b · ρb-Ryd? Rényi’s parking constant Figure 10: Comparison of b · ρb-Ryd? and b · ρb-Ryd∞ for 1 ≤ b ≤ 99. It can be shown that these two sequences approach different constants as b grows large lim b→∞ b · ρb-Ryd∞ = ∫ ∞ 0 exp [ −2 ∫ y 0 1− e−x x dx ] dy = 0.7475979202 . . . lim b→∞ b · ρb-Ryd? = 1. (3.13) 18 Ars Math. Contemp. 24 (2024) #P4.05 The constant appearing in the first limit is known as Rényi’s parking constant [57]. Both of these two limits are easier to understand in the context of irreversible deposition of k-mers. We deal with the k-mer deposition model in the following section where we revisit those limits. The calculation below, showing how to obtain the first limit in (3.13), and which we provide for completeness, appears in [32]. First note b∑ j=1 1− yj j = b∑ j=1 ∫ 1 y tj−1 dt = ∫ 1 y b∑ j=1 tj−1 dt = ∫ 1 y 1− tb 1− t dt = [ x = b(1− t) dx = −b dt ] = ∫ b(1−y) 0 1− (1− xb )b x dx, and therefore b · ρb-Ryd∞ = b ∫ 1 0 exp −2 b∑ j=1 1− yj j  dy = b ∫ 1 0 exp [ −2 ∫ b(1−y) 0 1− (1− xb )b x dx ] dy = [ ỹ = b(1− y) dỹ = −b dy ] = ∫ b 0 exp [ −2 ∫ ỹ 0 1− (1− xb )b x dx ] dỹ. The dominated convergence theorem now implies lim b→∞ b · ρb-Ryd∞ = ∫ ∞ 0 exp [ −2 ∫ y 0 1− e−x x dx ] dy = 0.7475979202 . . . Before we calculate the second limit in (3.13), we give a characterization of the value ρb-Ryd? in terms of a root of a certain polynomial. Compare this with the same results obtained by Došlić [20, discussion after Theorem 2.10] and Krapivsky–Luck [43, (3.4), (3.14) and (6.6)]. Theorem 3.9. The value ρb-Ryd? , at which the complexity of the Rydberg atom model with blockade range b, given in Theorem 3.5, attains its maximum, can be calculated as ρb-Ryd? = (1− z)(1− zb+1) 1 + b− bz − 2zb+1 − 2bzb+1 + zb+2 + 2bzb+2 , (3.14) where z is the unique root of the polynomial z2b+1 + · · ·+ zb+2 + zb+1 − 1, on the interval 0 < z < 1. Proof. We seek to find the density 12b+1 < ρ b-Ryd ? < 1 b+1 at which the complexity f = f b-Ryd in Theorem 3.5 attains its maximum. Again, we employ the Lagrangian function T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 19 method by setting L(ρ, z;λ) = ρ [ − ln 1− z 1− zb+1 − ( 1 ρ − (b+ 1) ) ln z ] − λ b∑ i=0 ( i+ b+ 1− 1 ρ ) zi = ρ ln 1− zb+1 1− z − (1− ρ(b+ 1)) ln z − λ b∑ i=0 ( i+ b+ 1− 1 ρ ) zi. The stationary points of this function solve the following system ln 1− zb+1 1− z + (b+ 1) ln z − λ ρ2 · 1− z b+1 1− z = 0 −ρ(b+ 1)z b 1− zb+1 + ρ 1− z − (1− ρ(b+ 1)) z − λ b∑ i=1 i ( i+ b+ 1− 1 ρ ) zi−1 = 0 b∑ i=0 ( i+ b+ 1− 1 ρ ) zi = 0. Using standard summation formulas, as in (3.12), it is possible to express ρ from the third equation as ρ = (1− z)(1− zb+1) 1 + b− bz − 2zb+1 − 2bzb+1 + zb+2 + 2bzb+2 . Plugging this into the second equation gives 0 = λ b∑ i=1 i ( i+ b+ 1− 1 ρ ) zi−1. From here, we conclude λ = 0. Finally, from the first equation we get λ = ρ2 1− z 1− zb+1 ln zb+1(1− zb+1) 1− z and, combining this with λ = 0, gives ln zb+1(1− zb+1) 1− z = 0, or zb+1(1− zb+1) = 1− z. We know from Theorem 3.5 that z 6= 1, so we can rewrite this equation as z2b+1 + · · ·+ zb+2 + zb+1 − 1 = 0. Clearly, there is a unique 0 < z < 1 solving this equation, and the corresponding ρb-Ryd? = (1− z)(1− zb+1) 1 + b− bz − 2zb+1 − 2bzb+1 + zb+2 + 2bzb+2 is the density at which the complexity in the Rydberg atom model with blockade range b is the largest. 20 Ars Math. Contemp. 24 (2024) #P4.05 The previous theorem can be used to give a proof of the second limit in (3.13). Corollary 3.10. lim b→∞ b · ρb-Ryd? = 1. Proof. Since 0 < z = z(b) < 1 solves the equation zb+1(1− zb+1) 1− z = z 2b+1 + · · ·+ zb+2 + zb+1 = 1 (3.15) it follows bz2b+1 < 1 < bzb+1 and therefore lim b→∞ z2b+1 = 0. Multiplying by z and taking square root we also get lim b→∞ zb+1 = 0. Finally, letting b→∞ in the identity zb+1(1− zb+1) = 1− z, gives lim b→∞ z = 1. Note that b · ρb-Ryd? = b(1− z)(1− zb+1) 1 + b(1− z)[1− 2zb+1]− 2zb+1 + zb+2 so in order to get limb→∞ b · ρb-Ryd? = 1, it suffices to show limb→∞ b(1− z) =∞. To see this, note that from (3.15) it follows (b+ 1) ln z = ln(1− z)− ln(1− zb+1) and hence lim b→∞ (b+ 1)(1− z) = lim b→∞ 1− z ln z · [ ln(1− z)− ln(1− zb+1) ] = −1 · [−∞− 0] = +∞ which completes the argument. 4 Complexity function of jammed configurations for irreversible de- position of k-mers It is easy to see that the Rydberg atom model with blockade range b is, up to scaling all densities by a factor b+1, equivalent to the irreversible deposition of k-mers model where k = b+1. As an immediate consequence of Theorem 3.5 we get the complexity of jammed configurations for irreversible deposition of k-mers. Corollary 4.1. For k ∈ N, k > 1, the complexity function of jammed configurations for irreversible deposition of k-mers is f(ρ) = { ρ k [ − ln 1−z 1−zk − ( k ρ − k ) ln z ] , if k2k−1 < ρ ≤ 1, 0, otherwise, T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 21 where z ≥ 0 is a real root of the polynomial k−1∑ i=0 ( i+ k − k ρ ) zi for which the expression f(ρ) is the largest. 2 3 3 5 4 7 5 9 1 2 1 ρ 0 0.05 0.1 0.3 lnw1 = lnw2 lnw3 lnw4 lnw5 lnw6 lnw7 lnw8 lnw9 lnw10 f (ρ ) k = 2 k = 3 k = 4 k = 5 k = 6 k = 7 k = 8 k = 9 k = 10 k = 11 Figure 11: The complexity function of jammed configurations for irreversible deposition of k-mers, for 2 ≤ k ≤ 11. Figure 11 shows the complexity function for all 2 ≤ k ≤ 11. Note that the support of the complexity function is now contained in the interval [1/2, 1]. In Figure 12 we compare the equilibrium density ρk-mer? and the jamming density ρ k-mer ∞ , for 2 ≤ k ≤ 100. In this model it is even more obvious that the Edwards hypothesis is violated. The limits of these two sequences as k grows large are lim k→∞ ρk-mer∞ = ∫ ∞ 0 exp [ −2 ∫ y 0 1− e−x x dx ] dy = 0.7475979202 . . . lim k→∞ ρk-mer? = 1. (4.1) Note that these limits are equivalent to those in (3.13). The convergence of jamming limits of k-mer deposition models (as k grows to infinity) to the Rényi’s parking constant is discussed in [44, §7.1] (see also [28, 32]). Clearly, the second limit from (4.1) follows from Corollary 3.10 as ρk-mer? = k · ρb-Ryd? for b = k − 1. Below, we provide a direct alternative proof of this fact. 22 Ars Math. Contemp. 24 (2024) #P4.05 Theorem 4.2. lim k→∞ ρk-mer? = 1. Proof. The quantity we are interested in, ρk-mer? , is equivalent to the quantity called the efficiency ε(k) in the context of packing Pk into Pn. It was shown in [20] that the efficiency is determined by the smallest singularity wk of the generating function Fk(1, y), i.e., by the smallest zero of its denominator. Hence we start by setting x = 1 into the rightmost expression in (2.4), Fk(1, y) = 1− yk 1− y − yk − y2k = 1−yk 1−y 1− yk 1−yk1−y . We rewrite its denominator as 1− qk(y), where qk(y) = qk 1−y k 1−y , and denote the smallest solution of equation qk(y) = 1 by wk. This equation has only one positive solution, since qk(0) = 0, qk(1) = k > 1 for large k and q′k(y) > 0 for all y > 0. Moreover, the same reasoning provides a better lower bound for wk, since qk( 12 ) = 2 (1−k)(1 − 2−k) < 1. Hence 1/2 < wk < 1. Consider now the expression ε(k) = ρk-mer? = k wkq′k(x) derived in [20]. First we rewrite q′k(wk) as q′k(x) = x k 1− xk 1− x [ 2k x − k x(1− xk) + 1 1− x ] . After plugging in x = wk, the term outside the brackets becomes equal to one, and by multiplying through by wk we arrive at wkq ′ k(wk) = ( 2− 1 1− wkk ) k + wk 1− wk . We are seeking upper bounds to the right-hand side. The first term is bounded from above by k, since the expression in parentheses cannot exceed one. It remains to bound the second term. As mentioned before, wk is the only positive solution of the equation 1− qk(x) = 0. We claim that, for a given (large) positive a, wk < 1 − ak for large enough k. So let us suppose otherwise, that for a given a > 0, wk > 1− ak is valid for all k. It means that the function 1− qk(x) has a positive value for x = 1− ak . By evaluating both sides, we obtain that ( 1− a k )k − ( 1− a k )2k < a k is valid for all k. This is a contradiction, since the left-hand side has a positive limit, e−a − e−2a > 0, while the right-hand side tends to zero as k tends to infinity. Hence, wk < 1 − ak for large enough k. Now the second term can be bounded from above by ak , and the whole expression for wkq′k(wk) is bounded from above by a+1 a k. Since a can be taken arbitrarily large, it means that the reciprocal value of wkq′k(wk), which is equal to our ρk-mer? , comes arbitrarily close to one, and our claim follows. T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 23 The convergence is quite slow, most likely logarithmic. We note another unusual thing in Figure 12. The equilibrium density ρk-mer? attains the minimum value for k = 9. The interpretation being that the polymers of length 9 pack the least efficiently of all polymers assuming the equilibrium model. This phenomenon was previously observed in [20]. 0 20 40 60 80 100 k 0.74 0.76 0.78 0.80 0.82 0.84 0.86 0.88 ρk-mer∞ ρk-mer? Rényi’s parking constant Figure 12: Comparison of ρk-mer? and ρ k-mer ∞ for 2 ≤ k ≤ 100. 5 Conclusions In this paper we have computed the complexity function (or configurational entropy) of jammed configurations of Rydberg atoms with a given blockade range on a one-dimensional lattice. We employed a purely combinatorial method which allowed us to compute the com- plexity function by solving a constrained optimization problem. Along the way we have explored and elucidated numerous connections between the considered problem and other models, such as, e.g., the random sequential adsorption and packings of blocks of a given length into one-dimensional lattices. In most cases, we have not followed those links very far. We believe that many interesting results could be obtained by deeper investigations of those connections. As an example, we mention here that explicit expressions for the number of maximal packings of given size from reference [20] could be directly translated into expressions for the number of jammed configurations of Rydberg atoms. By the same reasoning one can show that the total number of all jammed configurations of N Rydberg atoms with blockade range b on all one-dimensional lattices is given by (b+ 1)N+1. The methods employed here could be easily adapted for other one-dimensional struc- tures with low connectivity such as, e.g., cactus chains. Another class of promising struc- tures could be various simple graphs decorated by addition of certain number of vertices of degree one to each of their vertices. Similar problems were considered under various guises also for finite portions of rect- angular lattices, mostly for narrow strips of varying length. Among the best known prob- lems of this type are the so-called unfriendly seating arrangements. See [7, 30] for their history and some recent developments. To the same class belong the problems concerned with privacy, such as the ones considered in [39]. All cited references were concerned with one-dimensional lattices and/or narrow strips in the square grid, mostly with ladders. It would be interesting to consider those problems in finite portions of the regular hexagonal 24 Ars Math. Contemp. 24 (2024) #P4.05 lattice. Another interesting thing to do would be to study the behavior (and the difference) of ρ∞ and ρ? for different lattices/substrates. In other words, to investigate the difference between the jamming limit of dynamical models and the most probable densities in the equilibrium models. A drastic example is presented by the expected density of independent sets in stars: there are exactly two maximal independent sets in Sn = K1,n−1, one of them with size 1 and the other with size n−1. If both of them are equally probable, the expected size is n/2. Under dynamical model, however, the smaller one is much less probable than the bigger one, and the expected size is 1n + n−1 n (n − 1) = n − 2 + 2n . It would be interesting to know more about such differences and to know for which classes of graphs they are extremal. Our final remark is that the jammed configurations of Rydberg atoms with a given blockade range b are known as maximal b-independent sets in the language of graph theory. It might be worth investigating to what extent can similar problems be formulated also in terms of b-dominance in graphs. ORCID iDs Tomislav Došlić https://orcid.org/0000-0002-8326-513X Mate Puljiz https://orcid.org/0000-0003-0912-8345 Stjepan Šebek https://orcid.org/0000-0002-1802-1542 Josip Žubrinić https://orcid.org/0009-0002-6522-9554 References [1] M. C. Bartelt, J. W. Evans and M. L. Glasser, The car-parking limit of random sequential adsorption: Expansions in one dimension, J. Chem. Phys. 99 (1993), 1438–1439, doi:10.1063/ 1.465338, https://doi.org/10.1063/1.465338. [2] A. Baule, F. Morone, H. J. Herrmann and H. A. Makse, Edwards statistical mechanics for jammed granular matter, Rev. Modern Phys. 90 (2018), 015006, 58 pp., doi:10.1103/ RevModPhys.90.015006, https://doi.org/10.1103/RevModPhys.90.015006. [3] H. Bernien, S. Schwartz, A. Keesling, H. Levine, A. Omran, H. Pichler, S. Choi, A. S. Zibrov, M. Endres, M. Greiner, V. Vuletić and M. D. Lukin, Probing many-body dynamics on a 51- atom quantum simulator, Nature 551 (2017), 579–584, doi:10.1038/nature24622, https:// doi.org/10.1038/nature24622. [4] L. Berthier and G. Biroli, Theoretical perspective on the glass transition and amorphous ma- terials, Rev. Mod. Phys. 83 (2011), 587, doi:10.1103/RevModPhys.83.587, https://doi. org/10.1103/RevModPhys.83.587. [5] G. Biroli and R. Monasson, From inherent structures to pure states: Some simple remarks and examples, Europhys. Lett. 50 (2000), 155–161, doi:10.1209/epl/i2000-00248-2, https: //doi.org/10.1209/epl/i2000-00248-2. [6] B. Bonnier, D. Boyer and P. Viot, Pair correlation function in random sequential adsorp- tion processes, J. Phys. A: Math. Gen. 27 (1994), 3671, doi:10.1088/0305-4470/27/11/017, https://doi.org/10.1088/0305-4470/27/11/017. [7] H.-H. Chern, H.-K. Hwang and T.-H. Tsai, Random unfriendly seating arrangement in a dining table, Adv. in Appl. Math. 65 (2015), 38–64, doi:10.1016/j.aam.2015.01.002, https://doi. org/10.1016/j.aam.2015.01.002. T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 25 [8] E. R. Cohen and H. Reiss, Kinetics of reactant isolation. I. One-dimensional problems, J. Chem. Phys. 38 (1963), 680–691, doi:10.1063/1.1733723, https://doi.org/10.1063/ 1.1733723. [9] S. J. Cornell, K. Kaski and R. B. Stinchcombe, Domain scaling and glassy dynamics in a one- dimensional Kawasaki Ising model, Phys. Rev. B 44 (1991), 12263, doi:10.1103/PhysRevB.44. 12263, https://doi.org/10.1103/PhysRevB.44.12263. [10] A. Crisanti, F. Ritort, A. Rocco and M. Sellitto, Inherent structures and nonequilibrium dynamics of one-dimensional constrained kinetic models: a comparison study, J. Chem. Phys. 113 (2000), 10615–10634, doi:10.1063/1.1324994, https://doi.org/10.1063/ 1.1324994. [11] I. Csiszár and P. C. Shields, Information theory and statistics: a tutorial, Found. Trends Com- mun. Inf. Theory 1 (2004), 417–528, doi:10.1561/0100000004, https://doi.org/10. 1561/0100000004. [12] T. Cubel Liebisch, A. Reinhard, P. R. Berman and G. Raithel, Atom counting statistics in ensembles of interacting Rydberg atoms, Phys. Rev. Lett. 95 (2005), 253002, doi:10.1103/ PhysRevLett.95.253002, https://doi.org/10.1103/PhysRevLett.95.253002. [13] G. De Smedt, C. Godrèche and J.-M. Luck, Jamming, freezing and metastability in one-dimensional spin systems, Eur. Phys. J. B 27 (2002), 363–380, doi:10.1140/epjb/ e2002-00167-0, https://doi.org/10.1140/epjb/e2002-00167-0. [14] G. De Smedt, C. Godreche and J. Luck, Metastable states of the Ising chain with Kawasaki dynamics, Eur. Phys. J. B 32 (2003), 215–225, doi:10.1140/epjb/e2003-00091-9, https: //doi.org/10.1140/epjb/e2003-00091-9. [15] D. S. Dean and A. Lefevre, Steady state behavior of mechanically perturbed spin glasses and ferromagnets, Phys. Rev. E 64 (2001), 046110, doi:10.1103/PhysRevE.64.046110, https: //doi.org/10.1103/PhysRevE.64.046110. [16] D. S. Dean and A. Lefevre, Tapping spin glasses and ferromagnets on random graphs, Phys. Rev. Lett. 86 (2001), 5639, doi:10.1103/PhysRevLett.86.5639, https://doi.org/10. 1103/PhysRevLett.86.5639. [17] P. G. Debenedetti and F. H. Stillinger, Supercooled liquids and the glass transition, Nature 410 (2001), 259–267, doi:10.1038/35065704, https://doi.org/10.1038/35065704. [18] B. Derrida and E. Gardner, Metastable states of a spin glass chain at 0 temperature, J. Phys. France 47 (1986), 959–965, doi:10.1051/jphys:01986004706095900, https://doi.org/ 10.1051/jphys:01986004706095900. [19] T. Došlić, M. Puljiz, S. Šebek and J. Žubrinić, On a variant of Flory model, Discrete Appl. Math. 356 (2024), 269–292, doi:10.1016/j.dam.2024.06.011, https://doi.org/10.1016/j. dam.2024.06.011. [20] T. Došlić, Block allocation of a sequential resource, Ars Math. Contemp. 17 (2019), 79– 88, doi:10.26493/1855-3974.1508.f8c, https://doi.org/10.26493/1855-3974. 1508.f8c. [21] T. Došlić and I. Zubac, Counting maximal matchings in linear polymers, Ars Math. Contemp. 11 (2016), 255–276, doi:10.26493/1855-3974.851.167, https://doi.org/10.26493/ 1855-3974.851.167. [22] F. B. Dunning and T. C. Killian, Rydberg atoms: Giants of the atomic world, Scientia (2021), doi:10.33548/scientia679, https://doi.org/10.33548/SCIENTIA679. [23] Y. Elskens and H. L. Frisch, Aggregation kinetics for a one-dimensional zero-degree Kelvin model of spinodal decomposition, J. Stat. Phys. 48 (1987), 1243–1248, doi:10.1007/ BF01009543, https://doi.org/10.1007/BF01009543. 26 Ars Math. Contemp. 24 (2024) #P4.05 [24] J. W. Evans, Random and cooperative sequential adsorption, Rev. Mod. Phys. 65 (1993), 1281, doi:10.1103/RevModPhys.65.1281, https://doi.org/10.1103/RevModPhys.65. 1281. [25] P. Flajolet and R. Sedgewick, Analytic Combinatorics, Cambridge University Press, Cambridge, 2009, doi:10.1017/CBO9780511801655, https://doi.org/10.1017/ CBO9780511801655. [26] P. J. Flory, Intramolecular reaction between neighboring substituents of vinyl polymers, J. Am. Chem. Soc. 61 (1939), 1518–1521, doi:10.1021/ja01875a053, https://doi.org/ 10.1021/ja01875a053. [27] G. H. Fredrickson and H. C. Andersen, Kinetic Ising model of the glass transition, Phys. Rev. Lett. 53 (1984), 1244, doi:10.1103/PhysRevLett.53.1244, https://doi.org/10.1103/ PhysRevLett.53.1244. [28] H. D. Friedman, D. Rothman and J. K. MacKenzie, Problem 62-3, SIAM Review 6 (1964), 180–182, http://www.jstor.org/stable/2028090. [29] T. F. Gallagher, Rydberg Atoms, Cambridge Monographs on Atomic, Molecular and Chemi- cal Physics, Cambridge University Press, Cambridge, 1994, doi:10.1017/CBO9780511524530, https://doi.org/10.1017/CBO9780511524530. [30] K. Georgiou, E. Kranakis and D. Krizanc, Random maximal independent sets and the un- friendly theater seating arrangement problem, Discrete Math. 309 (2009), 5120–5129, doi: 10.1016/j.disc.2009.03.049, https://doi.org/10.1016/j.disc.2009.03.049. [31] L. Gerin, The Page-Rényi parking process, Electron. J. Combin. 22 (2015), Paper 4.4, 13 pp., doi:10.37236/5150, https://doi.org/10.37236/5150. [32] J. J. González, P. C. Hemmer and J. S. Høye, Cooperative effects in random sequential polymer reactions, Chem. Phys. 3 (1974), 228–238, doi:10.1016/0301-0104(74)80063-7, https:// doi.org/10.1016/0301-0104(74)80063-7. [33] W. Gotze and L. Sjogren, Relaxation processes in supercooled liquids, Rep. Prog. Phys. 55 (1992), 241–376, doi:10.1088/0034-4885/55/3/001, https://doi.org/10.1088/ 0034-4885/55/3/001. [34] C. S. Hofmann, G. Günter, H. Schempp, M. Robert-de Saint-Vincent, M. Gärttner, J. Ev- ers, S. Whitlock and M. Weidemüller, Sub-poissonian statistics of Rydberg-interacting dark- state polaritons, Phys. Rev. Lett. 110 (2013), 203601, doi:10.1103/PhysRevLett.110.203601, https://doi.org/10.1103/PhysRevLett.110.203601. [35] J. Jäckle and S. Eisinger, A hierarchically constrained kinetic Ising model, Z. Physik B - Con- densed Matter 84 (1991), 115–124, doi:10.1007/BF01453764, https://doi.org/10. 1007/BF01453764. [36] J. L. Jackson and E. W. Montroll, Free radical statistics, J. Chem. Phys. 28 (1958), 1101–1109, doi:10.1063/1.1744351, https://doi.org/10.1063/1.1744351. [37] D. Jaksch, J. I. Cirac, P. Zoller, S. L. Rolston, R. Côté and M. D. Lukin, Fast quantum gates for neutral atoms, Phys. Rev. Lett. 85 (2000), 2208, doi:10.1103/PhysRevLett.85.2208, https: //doi.org/10.1103/PhysRevLett.85.2208. [38] S. Kirkpatrick and D. Sherrington, Infinite-ranged models of spin-glasses, Phys. Rev. B 17 (1978), 4384, doi:10.1103/PhysRevB.17.4384, https://doi.org/10.1103/ PhysRevB.17.4384. [39] E. Kranakis and D. Krizanc, Maintaining privacy on a line, Theory Comput. Syst. 50 (2012), 147–157, doi:10.1007/s00224-011-9338-3, https://doi.org/10.1007/ s00224-011-9338-3. T. Došlić et al.: Complexity function of jammed configurations of Rydberg atoms 27 [40] P. L. Krapivsky, Kinetic models of a binary alloy at zero temperature, J. Stat. Phys. 74 (1994), 1211–1225, doi:10.1007/BF02188224, https://doi.org/10.1007/BF02188224. [41] P. L. Krapivsky, Dynamics of repulsion processes, J. Stat. Mech. Theory Exp. (2013), P06012, 27 pp., doi:10.1088/1742-5468/2013/06/p06012, https://doi.org/10.1088/ 1742-5468/2013/06/p06012. [42] P. L. Krapivsky, Large deviations in one-dimensional random sequential adsorption, Phys. Rev. E 102 (2020), 062108, 10 pp., doi:10.1103/physreve.102.062108, https://doi.org/10. 1103/physreve.102.062108. [43] P. L. Krapivsky and J. M. Luck, A renewal approach to configurational entropy in one di- mension, J. Phys. A 56 (2023), Paper No. 255001, 29 pp., doi:10.1088/1751-8121/acd5bd, https://doi.org/10.1088/1751-8121/acd5bd. [44] P. L. Krapivsky, S. Redner and E. Ben-Naim, A Kinetic View of Statistical Physics, Cam- bridge University Press, Cambridge, 2010, doi:10.1017/CBO9780511780516, https:// doi.org/10.1017/CBO9780511780516. [45] A. Lefèvre and D. S. Dean, Tapping thermodynamics of the one-dimensional Ising model, J. Phys. A 34 (2001), L213, doi:10.1088/0305-4470/34/14/101, https://doi.org/10. 1088/0305-4470/34/14/101. [46] J.-C. Lin and P. L. Taylor, Exact solution of a phase-separation model with conserved-order- parameter dynamics and arbitrary initial concentration, Phys. Rev. E 48 (1993), 4305, doi: 10.1103/PhysRevE.48.4305, https://doi.org/10.1103/PhysRevE.48.4305. [47] D. Lind and B. Marcus, An Introduction to Symbolic Dynamics and Coding, Cambridge Mathematical Library, Cambridge University Press, Cambridge, 2nd edition, 2021, doi: 10.1017/9781108899727, https://doi.org/10.1017/9781108899727. [48] J. K. Mackenzie, Sequential filling of a line by intervals placed at random and its application to linear adsorption, J. Chem. Phys. 37 (1962), 723–728, doi:10.1063/1.1733154, https: //doi.org/10.1063/1.1733154. [49] S. Masui, B. W. Southern and A. E. Jacobs, Metastable states of Ising spin glasses and ran- dom ferromagnets, Phys. Rev. B 39 (1989), 6925, doi:10.1103/PhysRevB.39.6925, https: //doi.org/10.1103/PhysRevB.39.6925. [50] M. Mézard, G. Parisi and M. A. Virasoro, Spin Glass Theory and Beyond, volume 9 of World Sci. Lect. Notes Phys., World Scientific, Singapore, 1987, doi:10.1142/0271, https://doi. org/10.1142/0271. [51] E. S. Page, The distribution of vacancies on a line, J. Roy. Statist. Soc. Ser. B 21 (1959), 364– 374, https://www.jstor.org/stable/2983806. [52] R. G. Palmer and H. L. Frisch, Low-and high-dimension limits of a phase separation model, J. Stat. Phys. 38 (1985), 867–872, doi:10.1007/BF01010420, https://doi.org/10. 1007/BF01010420. [53] D. K. Pickard and E. M. Tory, A critique of Weiner’s work on Palásti’s conjecture, J. Appl. Probab. 17 (1980), 880–884, doi:10.2307/3212986, https://doi.org/10.2307/ 3212986. [54] T. Pohl, E. Demler and M. D. Lukin, Dynamical crystallization in the dipole blockade of ultracold atoms, Phys. Rev. Lett. 104 (2010), 043002, doi:10.1103/PhysRevLett.104.043002, https://doi.org/10.1103/PhysRevLett.104.043002. [55] A. Prados and J. J. Brey, Analytical solution of a one-dimensional Ising model with zero-temperature dynamics, J. Phys. A 34 (2001), L453, doi:10.1088/0305-4470/34/33/103, https://doi.org/10.1088/0305-4470/34/33/103. 28 Ars Math. Contemp. 24 (2024) #P4.05 [56] V. Privman, Exact solution of a phase separation model with conserved order parameter dy- namics, Phys. Rev. Lett. 69 (1992), 3686, doi:10.1103/PhysRevLett.69.3686, https://doi. org/10.1103/PhysRevLett.69.3686. [57] A. Rényi, On a one-dimensional problem concerning random space filling, Publ. Math. Inst. Hungar. Acad. Sci. 3 (1958), 109–127. [58] M. Saffman, T. G. Walker and K. Mølmer, Quantum information with Rydberg atoms, Rev. Mod. Phys. 82 (2010), 2313, doi:10.1103/RevModPhys.82.2313, https://doi.org/10. 1103/RevModPhys.82.2313. [59] P. Sollich and M. R. Evans, Glassy time-scale divergence and anomalous coarsening in a kinet- ically constrained spin chain, Phys. Rev. Lett. 83 (1999), 3238, doi:10.1103/PhysRevLett.83. 3238, https://doi.org/10.1103/PhysRevLett.83.3238. [60] R. P. Stanley, Enumerative Combinatorics. Volume 1, volume 49 of Cambridge Studies in Advanced Mathematics, Cambridge University Press, Cambridge, 2nd edition, 2012, doi: 10.1017/CBO9781139058520, https://doi.org/10.1017/CBO9781139058520. [61] J. Talbot, G. Tarjus, P. R. Van Tassel and P. Viot, From car parking to pro- tein adsorption: an overview of sequential adsorption processes, Colloids Surf. A 165 (2000), 287–324, doi:10.1016/S0927-7757(99)00409-4, https://doi.org/10.1016/ S0927-7757(99)00409-4. [62] D. J. Thouless, P. W. Anderson and R. G. Palmer, Solution of ‘Solvable model of a spin glass’, Philos. Mag. 35 (1977), 593–601, doi:10.1080/14786437708235992, https://doi.org/ 10.1080/14786437708235992. [63] M. Viteau, P. Huillery, M. G. Bason, N. Malossi, D. Ciampini, O. Morsch, E. Arimondo, D. Comparat and P. Pillet, Cooperative excitation and many-body interactions in a cold Ryd- berg gas, Phys. Rev. Lett. 109 (2012), 053002, doi:10.1103/PhysRevLett.109.053002, https: //doi.org/10.1103/PhysRevLett.109.053002.