Zhang - Methods, Applications, and Challenges - 2020

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Received: 16 October 2020 Revised: 5 November 2020 Accepted: 6 November 2020

DOI: 10.1002/qua.26553

REVIEW

Global optimization of chemical cluster structures: Methods,


applications, and challenges

Jun Zhang | Vassiliki-Alexandra Glezakou

Physical Sciences Division, Pacific Northwest


National Laboratory, Richland, WA, USA Abstract
Chemical clusters are relevant to many applications in catalysis, separations, mate-
Correspondence
Jun Zhang, Physical Sciences Division, Pacific rials, and energy sciences. Experimentally, the structure of clusters is difficult to
Northwest National Laboratory, Richland, WA. determine, but it is very important in understanding their chemistry and properties.
Email: zhangjunqcc@gmail.com
Computational methods can be used to examine cluster structure, however finding
Funding information the most stable structure is not simple, particularly as the cluster size increases.
Office of Science, Grant/Award Number: DE-
AC05-76RL01830 Global optimization techniques have long been used to tackle the problem of the
most stable structure, but such approaches would have to look for a global minimum,
while sampling local minima over the whole potential energy surface as well. In this
review, the state-of-the-art theory of global optimization theory is summarized. First,
the definition, significance, relation to experiments, and a brief history of global opti-
mization is presented. We then discuss, in more detail, three versatile global optimi-
zation methods: the basin hopping, the artificial bee colony algorithm, and the
genetic algorithm. We close with some representative application examples of global
optimization of clusters since 2016 and the challenges, open questions and opportu-
nities in this field.

KEYWORDS

artificial bee colony, basin hopping, chemical clusters, genetic algorithm, global optimization,
potential energy surface

1 | I N T RO DU CT I O N

A chemical cluster is defined as an assembly of a few to several millions of structural units (SUs) such as atoms or molecules [1, 2]. This is a rather
general definition that covers various systems that can be completely distinct in chemical nature, like the examples in Figure 1(A)-(F). Chemical
clusters represent a bridge between a few atoms and bulk matter, exhibiting unique geometry and electronic structure [3–5], as well as physical
and chemical properties needed for catalysis [6–8], energy storage [9–11], new magnetic materials [12–14] and so on. Therefore, chemical clus-
ters have become the focus of numerous experimental and theoretical studies.
Unlike small molecules in the gas phase or crystals, the determination of cluster structures is rather nontrivial. Except for a few cases like C60,
the structural information of clusters can only be inferred indirectly from experimental measurements. For example, the infrared photodissociation
−1
spectroscopy of SO2−
4 ðH2 OÞ20 lacks the sharp band at about 3710 cm , suggesting that there is no dangling OH bond in the cluster (see
Figure 1) [15]. In direct atomic imaging experiments of Au55, no symmetric structures were observed, eliminating the possibility of icosahedral
structures [16]. With reliable structures, experimental phenomena can be interpreted and better understood and can be used to establish
structure–property relationships at the nanometer scale. Prediction of properties of clusters from first-principles requires efficient methods to
obtain reliable cluster structures.

Int J Quantum Chem. 2021;121:e26553. http://q-chem.org © 2020 Wiley Periodicals LLC 1 of 18


https://doi.org/10.1002/qua.26553
2 of 18 ZHANG AND GLEZAKOU

F I G U R E 1 Some typical
examples of chemical clusters:
(A) covalently bound atomic cluster
C60, (B) metal nanoparticle Ag32Cu6,
(C) micro-solvated SO2−4 ðH2 OÞ20 ,
(D) triethylphosphine-ligated cluster
Co6Te8(PEt3)6, (E) Au8supported on
oxygen-defected rutile (110) surface,
(F) an inverse double layer formed
by 500 1-palmitoyl-2-oleoyl-
glycero-3-phosphocholine (POPC)
molecules

Theoretical determination of small organic molecules has become a routine task for computational chemists. Given enough information
(bonding, stereochemistry, etc.), a good initial guess can be constructed. By choosing a suitable energy calculation method, like density functional
theory, one can calculate its potential energy as a function of the atomic coordinates within Born-Oppenheimer approximation, known as the
potential energy surface (PES). Starting with an initial guess structure, one can apply standard local optimization methods to locate a local mini-
mum (LM) on the PES. For clusters, however, this is an oversimplification of the problem, particularly as the size of the cluster grows. In a cluster,
there are numerous ways of arranging or packing its SUs that leads to many LMs on the PES. Among these LMs, the one with the lowest energy
is the global minimum (GM). The GM is believed to be the most abundant isomer in an ensemble of structures at low temperature and signifies
the optimal interaction patterns adopted by SUs in the cluster. Finding the GM of a cluster constitutes a global optimization (GO) problem. The
number of LMs grows exponentially with the size of the system [17, 18], thus a local optimization from a manual guess of a cluster structure
seldomly leads to the GM, and therefore more sophisticated methods are needed. The subject of this review is to discuss the state-of-the-art
methodologies for global optimization.
Before proceeding, an interesting question can be raised: “Does the GM of a cluster determined by a static optimization at 0 K correspond to
the structures inferred from experiments?” In practice, the answer is nontrivial. For a cluster whose low-lying isomers are not so energetically
close to the GM, it is often the best candidate that plays a role in experiments. For Au20, which was produced by laser vaporization of pure gold
targets, its lowest-lying LM is about 1.39 eV (32.05 kcal mol−1) above the GM. Properties calculated using the GM showed an excellent agree-
ment with photoelectron spectroscopy [19]. In other cases, LMs may come into play, especially when experiments are done at finite temperatures.
An example is water hexamer (H2O)6. Several accurate theoretical calculations suggest that the GM is the prism isomer [20–22], and the cage and
book isomer is about 0.09 and 0.21 kcal mol−1 higher in energy, respectively [20]. However, in 1996, only the cage isomer was observed in a ter-
ahertz laser vibration-rotation tunneling spectroscopy study [23]. In 2012, all three isomers were observed in a broadband rotational spectroscopy
study but the population was cage > prism > book [24]. Although the origin of this divergence is still unclear, we note that the cage and prism iso-
mer are almost energetically degenerate within “chemical accuracy” (1 kcal mol−1), thus the GM is qualified to represent (H2O)6, at least thermo-
dynamically. Cluster structures also depend on the way they are prepared. For example, electrospray ionization (ESI) tends to generate a larger
number of isomers than laser vaporization does. Sometimes isomers can be explicitly separated by means of, say, traveling-wave ion-mobility
spectrometry [25]. The reactions or physical chemical properties of these isomers can be studied. Even in these cases, the GO of clusters can pro-
duce not only the GM but also a set of LMs, i.e., a sampling over the PES, helping researchers use these isomers to draw a big picture of the whole
cluster PES and understand experimental observations. Thus, the GO of clusters is a meaningful and fundamental way to study clusters
theoretically.
The GO of clusters is much more challenging than the local optimization of chemical structures. Fortunately, with years of research and
improvement of computational hardware, the field has witnessed exciting and rapid development in the last three decades. In Table 1, some
selected applications of GO of clusters have been compiled (also see Refs [26, 27]). This is of course far from being complete: a simple search of
“global optimization chemical clusters” returns more than 50 000 publications since 2016. However, a general trend can be observed. Before
2005, people could perform GO only for small clusters with some cheap empirical potentials. Later, GO with first-principles methods were carried
−1=0= + 1
out for the first time for gas phase atomic clusters Li5− 7 (B3LYP) [28] and Ge12 − 20 (PBE) [29] in 2005, for surface-supported cluster Ag1 − 4,
6, 8, 10@MgO(100) (PBE) [30] in 2007, and for molecular cluster H(H2O)1 − 4 (B3LYP) [31] in 2010. Since 2016, clusters in GO studies go beyond

one- or two-component systems and become larger and more complicated, like (CH3OSO3H)x(H2SO4)y((CH3)2NH)z [32], Pt7H10CH3@α-Al2O3
(0001) [33], and Ag55@ZrO2/SiO2 [34]. These exciting changes enable GO of clusters to be applied to more realistic chemical problems.
ZHANG AND GLEZAKOU 3 of 18

TABLE 1 Selected applications of global optimization of clusters

Year System
1995–2000 (Morse)5 − 80 [35], LJ2 − 110 [36], (C6H6)2 − 15 (EP) [37], SC2 − 80 [38], (C6H6/C10H8/C14H10)2 − 6 (EP) [39], (H2O)2 − 21 (TIP4P) [40], LJ148
− +
− 309 [41], (NaCl)1 − 35Cl (CBM) [42], SinHm (MINDO/3) [43], H3O (H2O)1 − 20 (TIP4P) [44], (N2)13 − 55 (EP) [45]

2001–2005 (MgO)20 − 70 (CBM) [46], Na+/K+/Cs+(H2O)1 − 20 (CHARMM) [47], (Zn/Cd)3 − 125 (Gupta) [48], AgnCu45 − n (Gupta) [49], Si27 − 39 (NTB)
[50], (H2O)2 − 21 (TIP5P) [51], Li+/Ca2+(H2O)1 − 20 (CHARMM) [52], (SiO2)14 − 27 (EP) [53], (Al/Au/Pt)2 − 80 (VC) [54], (AlN)1 − 100 (EP)
− 1=0= + 1
[55] Li5− 7 (B3LYP) [28], Ge12 − 20 (PBE) [29],
+
2006–2010 C60(H2O)1 − 21 (H2O)1 − 21 (CHARMM) [56], B12− 25 (BP86) [57], Ag1 − 4, 6, 8, 10@MgO(100) (PBE) [30], (H2O)1 − 21 @graphite (TIPnP)

[58], Cu9 (PBE0) [59], Cs2 − 20 (SVWN) [60], (Al2O3)1 − 4AlO+ (B3LYP) [61], Au2 − 20 (SVWN) [62], Sn2− 15 (TPSS) [63], (C2H2)2 − 55 (EP)
[64], Aux Sy− (PBE) [65], Ag7 ðSR=DMSAÞ4− (PBE) [66], H(H2O)1 − 4 (B3LYP) [31]

2011–2015 (Li/Na/K)1, 2Si10 (PBE0) [67], (CH3OH)4 − 7 (B3LYPD) [68], Au6 − 12@MgO(100) (M06-L) [69], B21 (B3LYP) [70], Li+(CH3OH)1 − 20
(CHARMM) [71], (TiO2)2 − 13 (B3LYP) [72], C2 O2− 4 ð H2 O Þ 1−6 (BLYP) [73], Pd Ir
x y (PBE) [73], Pt3 − 6@CeO2 (111) (PBE) [74], Ir10 − 20
(PBE) [75], Tc4 − 20 (PBE) [76], SO2−
4 ðH2 OÞ3 − 50 (CHARMM) [77],(Au2S)1 − 8 (TPSS) [78], Si3 − 8C (BP86) [79], Au24@TiO2 (110) (PBE)
− 1=0= + 1 − 1=0= + 1
[80], Si20− 30 (B3PW91) [81], S3− 20 (B3PW9123) [82]
− 1=0= + 1 − 1=0
2016–2020 Pd2− 20 (B3LYP) [83], Pt12 − 46 (PBE) [84], PdxCoy (PBE) [85], ClO4− ðH2 OÞ3− 50 (CHARMM) [86], Na10 − 25 (B3LYP) [87], VGe1 − 19 (PBE)
− 1=0 − −
[88], RuxPty (PBE) [89], (NH3)20 (CHARMM) [90], ScSi20 (PBE) [91], Au60 (PBE) [92], TaB12 (PBE0) [93], (CO2)2 − 600 (CHARMM)
[94], (MgCl2)15(TiCl4)4 (PBE) [26], [bpy][SCN][SO2] (GFN) [95], (CH3OSO3H)x(H2SO4)y((CH3)2NH)z (CHARMM) [32],
ðI2 O5 Þx ðHIO3 Þy IO3− (CHARMM) [96], (Rh/Pt)43@ZrO2 (111) (PBE) [97], Pt7H10CH3@α-Al2O3 (0001) (PBE) [33], Ag55@ZrO2/SiO2
(PBE) [34]

Abbreviations: LJ, Lennard-Jones; CBM, Coulomb−Mayer–Born; EP, empirical potential; NTB, nonorthogonal tight-binding; SC, Sutton−Chen; VC,
Voter−Chen

This review provides a summary of the recent developments surrounding the GO of clusters. Several excellent reviews are available in this
field [27, 98, 99]. In the following sections, some of the most popular methods and recent applications of GO of clusters are summarized. We con-
clude with a brief perspective on the current challenges and future of global optimization approaches.

2 | METHODS

For local optimization, only the knowledge of the current point and those within a small neighboring area on the PES is needed. For GO, however,
the method has to explorer the whole PES, rendering GO a much more difficult task than local optimization. To enable sampling over the PES as
thoroughly as possible, the current mainstream approach is to use heuristic methods. In the context of this review, this terminology means that
instead of visiting the PES point by point, which would require astronomical amounts of computational resources and time, the PES is explored
from one LM basin to another. Some operations are used to escape from an LM basin, but success is not guaranteed. Therefore, a heuristic
method may or may not succeed in finding the true GM. Because no robust mathematical criteria for discriminating a GM exist, the GM obtained
from a GO can only be taken as a putative one, and chemical intuition is needed to answer the question: “Is there a better structure?”
The most essential part of a GO method is the ability to escape a LM basin, which defines the method and determines its average perfor-
mance. Currently, GO methods fall into two classes: (a) individual based ones that manipulate a single structure, and (b) population based ones
that examine a set of clusters, respectively, to explore the PES. The two classes of methods will be discussed in this section.
Before proceeding, we define the following terms that are often encountered: N is the number of atoms in the cluster; U is the potential
energy function; xgi is the coordinate of the ith cluster in the structure population in the gth cycle of GO; kB is the Boltzmann constant; and T is
the temperature.

2.1 | Potential energy surfaces

For chemical clusters, PES is a complicated hypersurface in the 3N-dimensional space. It is desirable to work with a transformed PES (see
Figure 2) rather than the original one U(x). A popular transformation is the smoothing procedure, which reduces the ruggedness of the PES: [100,
~ ðxÞ  minUðxÞ. Technically, this means that structures predicted in GO are always locally optimized. Smoothing is actually the Lamarckian
101] U
x
evolution in terms of a genetic algorithm. This removes the barriers to downhill movement toward a funnel [101], leading to more efficient GO
even though barriers between different funnels cannot be removed (see Figure 2). The most challenging part of GO is overcoming these barriers
smartly and efficiently. In some GO methods like stochastic surface walking [102], visited LM basins are filled using Gaussian functions (very simi-
lar to metadynamics [103]) to assist with removing barriers (see Figure 2).
4 of 18 ZHANG AND GLEZAKOU

F I G U R E 2 Transformation of the cluster PES. The black, red, and


blue curves are the original, smoothed, and Gaussian function-filled
PESs, respectively. The horizontal and vertical axes represent structure
and energy, respectively

F I G U R E 3 Illustrative PESs for


clusters dominated by either (A) long- or
(B) short-range interactions

The topological structure of a PES strongly depends on the interaction nature between the SUs in the cluster. Not strictly but inspiringly, the
interactions can be roughly divided into two classes: long-range and short-range ones. The former class includes Coulombic interactions, like
charge–charge or charge–dipole ones; the latter class includes covalent bonds, hydrogen bonds, and dispersion interactions. For clusters domi-
nated by long-range interactions, GO is relatively easier to accomplish than it is for clusters dominated by short-range interactions. This has been
observed in the GO of many model [42, 104, 105] and realistic clusters. For example, using the same GO method, it took only 1349 and 2798
steps to find the GMs of K+(H2O)20 and SO2−
4 ðH2 OÞ20 (charge-dipole interactions dominate), respectively, but 28 054 and 30 921 steps to find
those of (H2O)20 and CH4(H2O)20 (hydrogen bond and dispersion interactions dominate), respectively! [106]
A reasonable interpretation for this is that in a cluster dominated by short-range interactions, distant regions have relatively little impact on
each other, leading to many LMs that are similar in both energy and structure. A region with many such LMs on the PES is called a funnel (see
Figure 2). These LMs introduce barriers on the PES, making the PES very rugged; therefore, many GO methods are having a hard time escaping
from these funnels. For clusters dominated by long-range interactions, energy is more sensitive to local structural differences; therefore, the LMs
distribute themselves far away from each other on the PES. In the cases mentioned above, the presence of K+ or SO24 − strongly influences the
arrangement of H2O (with lone electron pairs on oxygen and oxygen–hydrogen bonds pointing toward K+ or SO2−
4 , respectively); other configura-
tions have much higher energies and are therefore more easily screened out during GO. For (H2O)20 or CH4(H2O)20, however, the clusters are
somewhat “homogeneous”—two clusters with a small number of different hydrogen bonds do not have significantly different stabilities, so GO
methods have to struggle somewhat with these close-lying structures (see Figure 3).

2.2 | Object functions

~ ðxÞ). For atomic clusters, it is logical to use Cartesian coordinates. For


The object function of the GO of clusters is the PES U(x) (or effectively, U
molecules, however, direct manipulation of Cartesian coordinates is inconvenient because the molecules may be unphysically destroyed; there-
fore, rigid-body coordinates are used for the whole SU instead. For example, in the six-component vector x = {α, β, γ, X, Y, Z}, the first 3 parameters
represent the three Euler rotation angles and the last 3 coordinates of the geometric center of the SU, respectively [106]. Some other choices are
available, but they do not exhibit an essential difference in GO performance [51]. Although the internal degrees of freedom (DOFs) of the SU are
ZHANG AND GLEZAKOU 5 of 18

frozen when only rigid-body coordinates are manipulated, this is usually a reasonable strategy for small SUs. Using (CH3OH)20 as an example, one
can consider each methanol as a rigid body and adjust the location and direction without taking into consideration the independent conformations
of methyl and hydroxyl groups. The latter can be relaxed during subsequent optimization via quantum chemical methods. Unfortunately, there are
indeed some cases where the internal DOFs must be treated explicitly, which will be discussed later.
The shape of U(x) strongly depends on the computational chemical methods, sometimes even changing the stability order of LMs. For exam-

ple, for two LMs of an ionic liquid (IL) cluster ½EMIM9 ½BF4 10 , the energy difference at GFN2-xTB [107] level of theory is +5.70 kcal mol−1, but at
B3LYP-D3 [108] level of theory is −10.83 kcal mol−1. Although the application of accurate methods in GO is always desired, the computational
cost is often prohibitive. It is often recommended that a cheap method like molecular mechanics (especially for organic molecular clusters), PM7,
or GFN2-xTB, is applied during the GO for organic molecular clusters. For inorganic or surface-supported clusters, however, a reliable first-
principles method should be used. Then, a set of good candidates (say, 100 of the best obtained structures) is further ranked with more accurate
first-principles methods like DLPNO-CCSD(T) [109]. A recent study suggested that using such a strategy, a reliable GM is very likely to be
found [110].

2.3 | Individual-based methods

Individual-based methods explore PES by changing a single cluster. Typical examples include basin hopping (BH) [36, 111–114], simulated
annealing (SA) [115], Monte-Carlo minimization (MCM) [116], and stochastic surface walking (SSW) [102]. Variations of these methods exist as
implemented by different researchers, but the basic procedure is similar: (1) Generate a starting structure (called seed). (2) Change the current
cluster structure using a preset criterion. (3) Decide whether to accept the new cluster or not. If not, the current cluster is reserved. (4) Repeat
(2) to (3) until the ceasing condition is satisfied. See Figure 4(A) for a visualization of BH.
The seed greatly affects the GO performance. A seed assumes the possible features of GM, which is both a challenge and a chance. On one
side, if one fails to provide a reasonable seed, GO may be trapped in a deep funnel, having to spend a very long time to find the true GM. On the
other hand, one can use prior knowledge of GMs of small-sized clusters to prepare a good seed to accelerate GO. For example, based on the
observation that intermediate-sized Lennard-Jones (LJ) clusters favor decahedral or icosahedral GMs, a GO based on decahedral or icosahedral
lattice search can efficiently find the GMs of large LJ clusters [117]. Later, a more reliable, unbiased GO of LJ clusters confirmed most of these
GMs [118]. More attractively, one can apply structural constraints on seeds. For example, in ABCluster [106, 119] and TGMin [120], a symmetric

structure generation algorithm based on Wyckoff positions [121] is implemented. For CoB18 , using 3 different seeds of C2v symmetry, the GM
can be located within 25 BH steps [120], while without this symmetry constraint, the GO took more than 5000 BH steps [122]. Using different
seeds, a multi-start GO can be performed, which is efficient for multi-funnel clusters like B20 [123].

F I G U R E 4 Summary of three
efficient and popular algorithms
discussed below: (A) basin
hopping, (B) artificial bee colony,
and (C) genetic algorithm
6 of 18 ZHANG AND GLEZAKOU

FIGURE 5 Typical moves applied in refining structures in BH

The essential difference between individual-based methods is the ways of generating a new cluster structure xg + 1 from the old one xg. In
BH, this is realized by moving all SUs: xg + 1 = xg + Δxg, where components of Δxg can be realized by a translation, a rotation, or a composite of
both (see Figure 5). When SUs are atoms or small molecules, independently moving every SU is sufficient; otherwise, block moves, which means
several SUs are moved together, are necessary for efficient GO [124]. Sometimes a chemically meaningful constraint, say, jΔxgj ≤ 2.5 Å or atomic
coordination number ≥ 3 is applied to prevent generating clusters that are too distorted.
Beyond this, more sophisticated forms of move exist. In SSW [102], structure change is realized by climbing. Starting from xg, a series of vec-
tors Nt are determined (by, say, biased dimer rotation method [125]) to drag the cluster to a high energy configuration xg, t. At the same time, a
Gaussian function G(x) is added to U(x) to fill up the LM basins, giving a transformed PES Ut(x) (see Figure 2). If the cluster relaxed on Ut(x) has
lower energy than xg, the climbing stops, and this cluster is relaxed on the original PES U(x) to give a new structure xg + 1.
After obtaining a new cluster xg + 1 from the old one xg, the method needs to determine whether to accept it or not. The most popular way is
the Metropolis acceptance criterion: [126] xg + 1 is accepted with the probability P(xg + 1j xg):

  ! !
  ~ xg + 1 − U
U ~ ðxg Þ
P xg + 1 jxg = min exp − ,1 ð1Þ
kB T

It implies that clusters of lower energy are always accepted, and the ones of higher energy are accepted with some probability. Acceptance of
higher energy structures is critical to overcoming the barriers between funnels, escaping from the current LM basin to a lower one. For an infinite
number of structure update steps at constant temperature T, Markov chain theory suggests that Equation (1) will lead to an ensemble of struc-
tures with Boltzmann distribution when some conditions are satisfied (like detailed balance, which guarantees an unbiased sampling) [127]. In BH,
SSW, and MCM, T keeps constant during GO. In SA, T is first increased to a very high value and is then decreased slowly. Since at elevated tem-
perature, it is more possible to accept clusters of high energy, the temperature increasing process is assumed to be able to increase the diversity
of sampling and jump out of the current funnel.

2.4 | Population-based methods

Population-based methods explore the PES by manipulating a set of clusters. They perform optimization by modeling natural processes such as
evolution and animal behaviors. The most classic one is the genetic algorithm (GA), which can be dated back to 1975 and might be the earliest
example of heuristic methods [128]. They have gained broad applications in engineering and industry. For GO of clusters, the most useful ones
include GA [129–139], artificial bee colony (ABC) [106, 119, 140], differential evolution (DE) [141–144], Kick algorithm [145–148], particle
swarming [149–153], and so on.
Generally, all population-based methods share the following steps: (1) Initialize a set of cluster structures (called population). (2) Update new
structures by applying operators on the individuals in the population. (3) Repeat (2) until the ceasing condition is satisfied. In this subsection, we
focus on two methods: ABC and GA (see Figure 4(B),(C)).
Unlike the individual-based methods, where the quality of seed structure affects the efficiency significantly (as shown by the example of

CoB18 mentioned above), population-based ones are less demanding in initialization. For clusters like ligated or surface-supported ones, some structural
constraints should be applied (for example, ligands should be directed toward the core; absorbates must stay above a specified surface) [140]. However,
for many clusters, the cluster can be simply initialized randomly, for example, components of x0 are given as random numbers in [0, L] where L is an esti-
mated size of the cluster [154, 155]. There are two ways to increase the population diversity: (1) Growing method—adding additional SUs randomly to
the GM or a LM of a smaller-sized cluster. (2) Chaos operator— [141, 156]initialize a new individual using a recurrence formula:
ZHANG AND GLEZAKOU 7 of 18

 0     
xi + 1 j = 4 x0i j 1− x0i j ð2Þ

The population size Npop can be adjusted according to the computational cost. Usually, 10–100 is sufficient [98, 119].

2.4.1 | Artificial bee Colony algorithm: ABCluster And NWPesSe

The ABC algorithm was proposed in 2005 in the computer science community by Karaboga and co-workers [157]. It is inspired by the foraging
strategy of honey bee colonies. A bee colony tries to find the best nectar as their food sources. It is observed that some bees, called employed
bees (EM), will first examine the surroundings, carrying information on nearby nectar candidates. The information can be shared through, for
example, waggle dances. Some bees, who just wait in the nest, called onlooker bees (OL), integrate information brought by EM bees. Finally, there
are also scout bees (SC),  5–10% of the total bee population, trying to search the surroundings for new nectars [158]. By modeling this strategy,
the ABC algorithm for GO of clusters has been proposed and is proven to be highly successful [106, 119, 140]. In this algorithm, three parameters
are used: population size SN, scout limit glimit, and maximum cycle number gmax. After initialization, the population is updated as follows:

1. EM bees: in cycle g, for each xgi ði = 1, , SNÞ, a new structure is generated by the trigonometric mutation operator: [159]

1 g       
x= xk1 + xgk2 + xgk3 + ðp2 −p1 Þ xgk1 −xgk2 + ðp3 −p2 Þ xgk2 −xgk3 + ðp1 − p3 Þ xgk3 −xgk1 ð3Þ
3

  
~ g 
U xkm 
pm =          ðm = 1, 2,3Þ ð4Þ
~ xg  + U
U ~ xg  + U
~ xg 
k1 k2 k3

where km are random integers in [1, SN], satisfying k1 ≠ k2 ≠ k3. It is highly efficient in exchanging information between individuals [106,
119, 141, 159]. Next, the greedy selection scheme is used to decide if the generated one should be accepted:

(  
~ xg ≤ U
xgi , U ~ ðxÞ
xgi + 1 = i
 g ð5Þ
x, U ~ x >U ~ ðxÞ
i

Because the large size of a population can offer structural diversity, it is unnecessary to employ a Metropolis acceptance criterion as in
individual-based methods to keep clusters of higher energy. The EM bees introduce multiple interactions between individuals.

2. OL bees: this step tries to do a search based on “good” clusters (low energy ones) in the population. The good clusters can be selected in two
ways: the roulette wheel and the tournament selection [98]. In the former one, the “good” cluster is selected with a probability proportional to
its fitness; in the second one, Nplayer individuals are randomly picked up and the best one among them is selected. Obviously, the roulette
wheel selection always biases toward the cluster xgbest that has the lowest energy in the population. Therefore, the tournament selection is pre-
ferred. The OL operator generates a new structure by:

8  
>
< xgk + F xgk1 + xgk2 −xgk3 − xgk4 , η ≤ 0:5
x=   ð6Þ
>
: xgbest + F xgk + xgk − xgk −xgk , η > 0:5
1 2 3 4

Here, η and F are random numbers in [0, 1), km (m = 1, 2, 3, 4) are random integers in [1, SN] that satisfy k1 ≠ k2 ≠ k3 ≠ k4, and xgk is the clus-
ter obtained from a tournament selection. The two expressions are denoted by “ABC/current/2” and “ABC/best/2”, respectively, which can
improve the performance of the ABC algorithm [160]. The OL bees is a positive feedback mechanism in the algorithm.

3. SC bees: in this step, each xgi is examined. If it is not updated in the last glimit generations, it will be discarded and replaced by a random struc-
ture. This is a negative feedback mechanism.
8 of 18 ZHANG AND GLEZAKOU

(1)–(3) are repeated for gmax generations (see Figure 4(B)), which is the algorithm implemented in ABCluster. In NWPEsSe [140], which aims
at nanosized systems, in each cycle, only one of the EM, OL, and SC bees is randomly applied to the population. The generated cluster is
appended to the population instead of replacing one. This is a compromise between computational cost and efficiency.
Note that besides multiple interactions and positive and negative feedback, the random numbers in the working equations introduce fluctua-
tions in the algorithm. These four features make the ABC algorithm a self-organizing [161] and self-learning [140] one. The combination of EM,
OL, and SC bees guarantee that the algorithm will not stagnate and get trapped in a LM basin.

2.4.2 | Genetic algorithm

GA is probably the most famous heuristic method. It is based on the theory of evolution. To survive in natural selection, individuals have to exhibit
high fitness; thus, they must develop strategies to maintain the diversity of genes within their populations, two familiar ways of which are chro-
mosomal crossover during mating and mutation in-place. After many generations of natural selection, the individuals with low fitness will be elimi-
nated, and the overall fitness of the entire population gets increased. In the context of GO of clusters, GA is implemented by introducing the
mating, crossover, and mutation operators to generate new structures and by using “natural selection” to improve the population. In early studies
[129, 162], clusters were encoded to binary strings, and crossover and mutation were applied using bit operations. Since Zeiri's work [98], GA can
be directly applied to Cartesian coordinates. A typical GA procedure is shown in Figure 4(C) [98]. Four parameters are needed: population size
Npop, mating times Nmat, mutation probability pmut, and maximum cycle number gmax. After initialization of the population:

1. Calculate the fitness fi. For each individual xgi , its energy is first normalized to [0, 1]:

 
~ xg − U
U ~ min
ρi = i
ð7Þ
U ~ min
~ max − U

~ max and U
where U ~ min is the maximum and minimum energy of clusters in the current population, respectively. fi can be calculated with different fit-
ness functions [98], for example, in applications of GA on (MgO)n [46] and 15MgCl2/4TiCl4 [26] clusters, a hyperbolic tangent and exponential
function was used, respectively:

1 −tanhð2ρi −1Þ
fi = ð8Þ
2

f i = expð −3ρi Þ ð9Þ

2. Mating and crossover. Two clusters xg1 and xg2 are picked using the roulette wheel or the tournament selection for mating. The new structures
(called offspring) are generated using a crossover operator: “cut-and-splice” (see Figure 6) [130]. For atomic clusters, one can apply one/two-
point crossover: both individuals are cut along one/two surface and are spliced with another/central fragment from each other [98]. Technical
realization of this must preserve the correct stoichiometry and basic skeleton [26]. Such mating and crossover is carried out for Nmat times.

3. Mutation. The mating and crossover only recombine the existing structural motifs, but mutation can introduce new structural motifs to avoid
population stagnation. This is very similar to SC bees in the ABC algorithm. Mutation operators are performed on all offspring with a probabil-
ity pmut. There are a lot of forms of mutation operators, like rotating a part of the cluster, relocating a part of the clusters, and exchanging two
atoms of different types (see Figure 6). The exchange operator is highly efficient for multi-component atomic clusters.
4. Selection. The Npop most stable clusters from the current population together with the offspring are kept, while others are eliminated. This is
again a greedy selection Scheme.

(1)–(4) are repeated for gmax generations (see Figure 4(C)).


Using chemically meaningful crossover and mutation operators for specific systems, one can carry out efficient GA.

2.5 | Implementation and software

To make these various GO methods available for chemists, it is highly desirable to implement them in portable and easily accessible codes. Several
codes are available, like PDECO [141], GEGA [28], GMin [163], TGMin [120], AUTOMATON [137], and CALYPSO [153]. They implement the DE,
ZHANG AND GLEZAKOU 9 of 18

F I G U R E 6 Examples of mating and


mutation operators in GA

GA, or BH method. The ABC algorithm has been implemented in the software ABCluster [106, 119] and NWPEsSe [140]. Both are highly user-
friendly, well-documented, and powerful. They have shown excellent performance in the GO of gas phase, ligated, surface-supported, and assem-
ble of atomic and molecular clusters. Some realistic applications are given in the subsequent section.

3 | APPLICATIONS

The GO of clusters is critical to chemists for structure determination and property calculation. Only reliable structures can guarantee the meaning-
fulness of the subsequent studies. Because of its fundamental importance, GO has broad applications in all chemical fields. Due to limited space,
only some recent applications are discussed.

3.1 | Atmospheric chemical processes

New-particle formation (NPF) [164] is defined as the formation of clusters of atmospherically related molecules such as water, nitric acid, ammo-
nia (NH3), sulfuric acid (H2SO4), sulfamic acid (H2NSO3H), dimethylammonia ((CH3)2NH), glycolic acid, lactic acid, malonic acid, iodine pentoxide,
iodine acid, and methanol, to name a few. NPF plays a critical role in many global climate phenomena and is the source of more than 50% of aero-
sol particles [165], participates in cloud condensation [166], and has an impact on the distribution of air pollutants [167]. Understanding NPF can
significantly improve global climate modeling [168] and decrease negative health effects caused by pollutants and ultrafine particles [169].
Molecular-level details of NPF can be studied by looking at the formation free energies of clusters of n molecules (ΔGn) calculated via computa-
tional quantum chemistry. The evaporation rate γn + m ! n can be derived from ΔGn, ΔGm, and ΔGn + m [170]. The time evolution of molecular clus-
ter distributions can then be obtained with the atmospheric cluster dynamic code (ACDC), which solves a set of birth-death equations [171].
However, calculations of ΔGn need GMs of either heterogeneous or homogeneous clusters containing the molecules mentioned above. These
molecules contain different types and numbers of interaction sites, and manual enumeration of possible configurations becomes a daunting task
when n > 2. This has been one of the most significant challenges in such studies before 2017.
The development of GO as implemented in ABCluster, has eliminated this obstacle to a large extent, making a series of productive NPF stud-
ies possible [32, 96, 167, 172–197] and revealing valuable insights into atmospheric chemistry (see Figure 7 for some examples). ABCluster has
been regarded as a standard tool in this field [198–200]. In a work by Zhang, Francisco, and Zeng [176], the GMs of
(H2NSO3H)x(H2SO4)y((CH3)2NH)z (0 ≤ x + y ≤ 3, 0 ≤ z ≤ 2) were searched with ABCluster and were subsequently used in quantum chemical and
ACDC simulations. The dominant loss pathway of the major air pollutant sulfur trioxide (SO3) is believed to be its reaction with water that pro-
duces H2SO4 [201]. Zhang and co-workers [176] suggested a new loss pathway—the reaction of SO3 with NH3 to produce H2NSO3H. The latter
reaction can be self- and auto-catalyzed by NH3 and H2NSO3H, respectively. Moreover, H2NSO3H was found to directly participate in NPF from
H2SO4 and (CH3)2NH and to increase the rate of NPF by a factor of 2. Therefore, in dry and NH3-rich (heavily polluted) air, SO3 could be primarily
depleted by NH3 (instead of water), with the product H2NSO3H accelerating NPF and affecting the local air quality.

3.2 | Discovery of novel structures

For clusters in the gas phase, accurate spectroscopy information like photoelectron spectroscopy can be obtained. Therefore, a productive
method is to perform GO of the clusters and then compare the calculated and experimental spectra to determine the cluster structures. Some
example structures are shown in Figure 8.
Fifty-five is considered to be a geometric magic number for atomic clusters because this number of particles can form structures of icosahe-
dral symmetry, which is presumed to be highly stable. Several groups [202, 203] applied GO to M55 to see if this is true for metal clusters. It turns
10 of 18 ZHANG AND GLEZAKOU

F I G U R E 7 Example GMs of molecular clusters obtained with ABCluster: [106, 119] (A) (HNO3)3(NH3)3(H2O)2 [167], (B) (H2NSO3H)
 
(H2SO4)2((CH3)2NH)3 [176], (C) ðI2 O5 Þ4 IO3− 3 ðH2 OÞ10 [96], (D) (Gly)2(H2SO4)(NH3)2 [184]

− −
F I G U R E 8 Example GMs of atomic clusters obtained with (A) NWPEsSe, (B, C) TGMin, and (D) GEGA: (A) Au55 [140], (B) B42 [228], (C) TaB20
+
[231], (D) Li7 ðBHÞ5 [236]

out that the GMs of Ni55, Cu55, and Ag55 do have icosahedral symmetries, but the GMs of Pd55, Pt55, and Au55 prefer disordered structures, which

results from the competition between dense packing and relativistic effects [204]. The gold cluster anions were examined for Au21 −26,42− 50
[205–207], all of which were found to be disordered. Nevertheless, in 2019, Gong, Wang, and Zeng applied photoelectron spectroscopy and a
large population of BH to Aun− and confirmed that Au60

has a high icosahedral symmetry [92].
Yang et al. used ABCluster to study metal-doped silicon clusters MSiqn (q = 0, − 1, n ≤ 20, M = Sc, Y, Ti, Zr, and some lanthanides) [91,
116, 208–220]. The computed photoelectron spectroscopies using GMs agree well with experimental ones. Although the shapes of GMs strongly
depend on n, q, and M, a common observation is that metal atoms tend to be surface-attached, linked, and completely encapsulated as n becomes
larger. Wang and Zeng studied some gold alloy clusters like Aux Aly− (x = 2, y = 3 − 11 or x + y = 7, 8) [221, 222] and AuBi4−−8 [223], all of which rev-
ealed multiple close-lying polyhedral structures.
Boron clusters have attracted intense attention because of the electron deficiency of boron and because their GMs possess unique features.
For medium-sized clusters, a series of joint photoelectron spectroscopy and BH studies by Wang and Li suggested several interesting structures,

including B26,30,35 −38,40− 42 (with one or two hexagonal vacancies) [224–229]. For metal-doped boron clusters, the GMs exhibit much higher sym-
metry than MSin− does. A few examples include metal-centered [230] ReB8,9
− −
, tubular [231, 232] MnB16 −
, TaB20 , half-sandwich [93, 233, 234]
SmB6− , PrB7− , TaB12

, and inverse sandwich [235] La2 B8− , Pr2 B8− . Using different variants of GA, Li7 E5+ (E = BH, C, Si, Ge) clusters were found to be
−1=0
star shaped [236–239] and aromatic. For some clusters like ðMg=Al=ZrÞB10 −20 [240–242], both anionic and neutral forms were studied. The struc-
tures, which were obtained with CALYPSO, revealed that except for a few cases, anionic and neutral GMs have very similar geometries.

3.3 | Exploring new materials

The importance of searching for new materials can never be overemphasized. Two forms of GO are useful in this field. The first form is GO of
solid structures, i.e., looking for crystal and amorphous structures depending on the components of the materials, which can be done with some
programs like CALYPSO [153] or USPEX [243]. This topic is beyond the scope of this review. The second form is GO of clusters. The second
method is much faster than the former one and provides important chemical insights at the atomic level.
ZHANG AND GLEZAKOU 11 of 18

For example, to study a non-conjugated photoluminescent polymer for bioimaging [244], an ethyleneimine tetramer was generated using
ABCluster to model polymer aggregation. Quantum chemical calculations implied better-conjugated electronic structure and increased rigidity of
N-heteroatoms in the original polymer after aggregation, making it a promising nanoprobe in bioimaging. In a study of lithium-sulfur batteries
(LSBs) [245], GMs were searched for polysulfide Li2Sn on C2N nanosheets using ABCluster. Subsequent calculations revealed that polysulfides
can be quite stable on C2N without dissolution, indicating that C2N is a good cathode material candidate for LSB. ABCluster was also used to
investigate some clusters of metallic hydrogen energy materials [246, 247].
Although a cluster sometimes does not demonstrate outstanding properties on its own, it is possible to ligate the cluster and tune its proper-
ties [248, 249]. The GMs of the bare and ligated clusters can have very different structures. For example, while Au28 + has a triangle-like structure,
 2 +  2 +
the gold core of Au8 PðCH3 Þ3 5 is hexagon-like, and that of Au8 PðCH3 Þ3 8 is cubic (see Figure 9(A)-(C)) [140]. This implies that ligation can be
used to tune the size of the gold clusters, which can be employed to control the inter-distance between graphene flakes in solution using separa-
tion techniques. Another example of a ligated cluster is Co6Te8(PEt3)6 (see Figure 1(D)). Although ligation does not change the core structure, it
does stabilize this electron-rich cluster. When ligation of C60 (which is a good electron acceptor) is successful, a promising energy material can be
obtained [250]. The GO from NWPEsSe suggested that the coordination sphere of Co6Te8(PEt3)6 can hold up to 20 C60, leading to a 3 nm system
(see Figure 9(D)) [140]. This is a large and challenging system for GO.

3.4 | Calculations of thermodynamic properties

Some thermodynamic properties of compounds, like solvation energies or electric permittivity, result from the collective motion of many mole-
cules. Calculating these properties from a single molecule is problematic and therefore must be done with clusters. For example, calculating hydra-
tion energies of charged species by applying implicit solvation models on a single molecule exhibits systematic errors, and a thermodynamic cycle
model involving solvent and solvent–solute clusters provides much better results [251]. In the latter case, an efficient global optimizer that can
obtain the GMs of explicitly solvated clusters is much preferred.
Malloum studied a series of ammonia clusters [90, 252, 253], including (NH3)20,30,40,50 and H+(NH3)18,20,25,30,40,50. The infrared (IR) spectra
computed using these GMs agree quite well with experimental spectra. Using 50-mers, the solvation free energy of the proton in ammonia was
calculated to be −277 kcal mol−1 at the M06-2X/6–31++g(d,p) level of theory. Malloum also investigated other molecular clusters [254–256].
Huang found that the real and imaginary part of the effective permittivities of the pyridine (C5H5N)–ethanol (C2H5OH) mixture are larger than
those of their pure components at some concentrations [257]. Calculations on the GMs of (C5H5N)n(C2H5OH)m (n = 0 − 4, m = 0 − 6) suggested
that this difference is related to the dipole moments of the mixture. Cao generated structures of hydrated lanthanide(III) motexafins [258]. It was
observed that only when explicit water molecules exist can the calculated UV–Vis absorption spectrum match the experimental one. All clusters
mentioned in this paragraph were studied with ABCluster.
Recently, Zhang et al. observed the occurrence of IL clusters ½EMIMn ½BF4 n−+ 1 (n = 1 − 9) in ESI mass spectra (MS) (see Figure 10(A)) [259].
Using NWPEsSe, the GMs of ½EMIMn ½BF4 n−+ 1 were searched, and calculations using these GMs indicated that ½EMIM2 ½BF4 3− has the highest
stability and should be the most abundant cluster in ESI-MS, in excellent agreement with the experimental data. The GMs also revealed that the
clusters change from packed to liquid-like disordered structures, as indicated by changes in the Coulombic interaction strength (see Figure 10
(B)) [259].

3.5 | Revealing catalysis mechanisms

GO of clusters has been used to reveal homogeneous catalytic mechanisms. In 2016, Shaik and Schwarz observed that a metal oxide cluster,
Al2 Mg2 O5+ , can activate methane [260]. Its structure, determined by ABCluster, revealed that the cluster can activate methane through both
proton-coupled electron transfer and hydrogen-atom transfer mechanisms. The first step in uncovering the catalytic roles of the nanoparticles
0= + 1
Pt24 [261] and VB4,5 [262, 263] was also made via ABCluster structure determination. To model liquid-phase reactions, ABCluster has been used
to construct clusters of reaction intermediates and solvent molecules [264–266], for example, in the study of electrochemically catalyzed New-
man−Kwart rearrangement [267].
GO has also been widely applied in heterogeneous catalysis. By using GA to search the cluster 15MgCl2/4TiCl4, Taniike proposed a model for
the Ziegler−Natta catalyst [26]. Sautet and Alexandrova applied BH to investigate the configurations of the reaction complex adsorbed on
α-Al2O3 (0001) [33]. In the conversion of ethanol, an amorphous catalyst—ZrO2/SiO2-supported Ag—exhibited good performance. NWPEsSe was
used to carry out GO of silver films and nanoparticles on ZrO2/SiO2 [34]. The two different dispersion forms of silver were theoretically found to
promote conversion of ethanol to 1,3-buitadiene and ethylene, respectively, which was in agreement with experimental results. NWPEsSe was
also applied in the investigation of the PES of oxygen-defected rutile TiO2(110) surface-supported Au8 [140]. In the GM, two Ti–Au covalent
bonds are formed, and Au8 takes a different geometry from its GM in the gas phase. In addition, the GM and other LMs are very close-lying,
12 of 18 ZHANG AND GLEZAKOU

indicating a flat PES. Thus, upon absorption, Au8 becomes a charge reservoir and liquid-like. This behavior was believed to play an important role
in its heterogeneous catalysis [268]. See Figure 11 for examples of the surface-supported clusters mentioned above.

4 | C HA L L E N G E S A N D P E RS P E C T I V E

While GO of clusters has been increasingly applied in realistic chemical problems, several challenges persist in this field, some of which are dis-
cussed in this section.

4.1 | Manipulating internal degrees of freedoms

Although treating small SUs in a cluster as rigid bodies is often a reasonable strategy, for large SUs that have several internal DOFs, this strategy
could be completely misleading. For n − C20H41OH clusters, the stable conformation of the long alkyl chain of n − C20H41OH may be all-trans,
hairpin, or some other form [269]; these different structures are separated by multiple barriers on the chain's PES. Internal DOFs of SUs must be
explicitly sampled, which is a problem when employing a conformation search (CS) [270]. For (n − C20H41OH)20 and similar cases, the CS can be
performed first, then clusters can be constructed using these conformers during the GO. This strategy has been used in some cases [198].
Unfortunately, in complicated cases, GO and CS are inseparable. A typical example is the ligated gold cluster Aun Lqm . In many cases, only some
of the Au atoms form a core, and the remaining Au (called exo) form various types of “staple” motifs with L, such as AukLk + 1 [271, 272]. In these
staple structures, L can mono- or multi-dentate ligands like bis(diphenylphosphino)propane. These motifs must be manipulated carefully. For
Aun Lqm , the numbers of core and exo gold atoms and the possible numbers and types of staple motifs are hard to predict and must be determined
during GO. Even when the information is known (for example, from experiments), the external and internal DOFs of these motifs should be
adjusted to avoid clashes with themselves and others, while their anchoring atoms like sulfur or phosphorous can coordinate with surface Au on
the core, the structure of which is also determined in GO. Therefore, Aun Lqm raises a grand challenge to GO. Although BH, SA, and GA have been
successfully applied for small systems like Au20(SCH3)16 [273], generally no robust and efficient methods exist for such systems.

4.2 | Applications of machine learning to global optimization

As clusters become larger, the number of energy evaluations for reliable GO increases exponentially. This is a serious bottleneck in GO, especially

for inorganic and surface-supported clusters for which no robust and reliable force-field-like methods exist. For example, Au60 was discovered by
examining more than 10 000 isomers at the spin-orbital PBE0 level of theory in a BH GO [92]. For [Co6Te8(PEt3)6][C60]20/21 (see Figure 9), even a
local optimization at the GFN2-xTB level of theory would take almost one day on ordinary hardware. Besides continuing to develop new quantum
chemical methods, machine learning (ML) is becoming a promising way to accelerate GO and solve this dilemma. ML has been used in both the
energy evaluation and structure generation steps of GO.
ML potentials (MLPs) are becoming more popular in computational chemistry [274]. The large number of parameters in MLPs impart suffi-
cient flexibility to describe the complicated PESs at the cost of force fields, but with an accuracy close to first-principles methods. GO could bene-
fit tremendously from MLPs for specific elements. Two examples are GO of gold [275] and aluminum [276] atomic clusters. Their PESs are fitted
to different forms of MLPs (like the Gaussian process or deep neural networks), enabling fast GO of up to 100 atoms with an accuracy close to

F I G U R E 9 Example GMs of
ligated clusters obtained with
NWPEsSe: (A) Au28 + ,
 2 +
(B) Au8 PðCH3 Þ3 5 ,
 2 +
(C) Au8 PðCH3 Þ3 8 ,
(D) [Co6Te8(PEt3)6][C60]20
ZHANG AND GLEZAKOU 13 of 18

FIGURE 10 The (A) ESI-MS and (B) GMs of ½EMIMn ½BF4 n−+ 1 . Figure adapted from Ref. 259

F I G U R E 1 1 Example GMs of surface-supported clusters obtained with (A) PGOPT and (B, C) NWPEsSe: (A) Pt7H10CH3@α-Al2O3 (0001)
(PBE) [33], (B) Ag55@ZrO2/SiO2 (PBE) [34], (C) Au8@oxygen-defected rutile TiO2(110) [140]

that of first-principles methods. There are several forms of MLPs in this field, like Behler–Parrinello symmetry functions [277], deep potentials
[278], and embedded atom neural network potetial [279]. There is no consensus regarding which is the most suitable approach for GO of cluster
structures. In addition, fitting reliable MLPs is a nontrivial task. How to construct a compact training set with sufficient structural diversity is still
an open question. Sometimes, instead of fitting the MLP before GO, it is fitted on the fly using the clusters generated during GO [280–282]. Such
crude MLPs are used to screen the clusters that most likely have high energies before performing expensive first-principles calculations.
Some sophisticated ML models have been used to guide generation of new structures. Reinforcement learning was applied in the reconstruc-
tion of material structures [283, 284]. In this scheme, a policy is defined as a subroutine P that determines where a new atom is added based on
the current state (atoms already constructed). When the added atom results in a stable structure, a high reward is given to P. Thus, P is trained
using the structures generated during GO to always make determinations for a high reward. A good P is very helpful in predicting complicated
structures, like oxygen on Ag(111) [283]. These ML schemes are quite novel and probably require some modifications when applied to clusters.
Nevertheless, their potential use in improving the performance of cluster GO is promising.

5 | OUTLOOK

In this review, we present a brief history as well as an overview of contemporary state-of-the-art methods for global optimization of cluster struc-
tures. Thanks to the development of user-friendly software, global optimization of extended clusters can be efficiently undertaken by non-experts
in this field, becoming a powerful tool. From recent representative applications, we can see that the GO of chemical cluster structures plays an
important role in understanding a broad range of chemical phenomena and determination of chemical properties. Nevertheless, there are still
unsolved, open questions in this field calling for more elegant algorithms that can also benefit from the recent surge in data science approaches.
The traditional heuristic algorithms like BH, ABC, and GA can be organically combined with novel methods like ML to treat overcome difficulties
associated with large sizes of systems. We hope that the ideas from this topic can also inspire related fields, for example, ligand docking, predic-
tion of crystal structures and amorphous materials. We believe that the global optimization of cluster structures is a growing field with potential
of broader applications.
14 of 18 ZHANG AND GLEZAKOU

AUTHOR CONTRIBUTIONS
Jun Zhang: Inverstigation; formal analysis; methodology; writing-original draft; writing-review and editing. Vassiliki-Alexandra Glezakou: Funding
acquisition; project administration; supervision; writing-review and editing.

ORCID
Jun Zhang https://orcid.org/0000-0001-9093-9430

RE FE R ENC E S
[1] R. L. Johnston, Atomic and molecular clusters, CRC Press, London 2002.
[2] T. Fehlner, J.-F. Halet, J.-Y. Saillard, Molecular clusters: a bridge to solid-state chemistry, Cambridge University Press, Cambridge 2007.
[3] M. Ha, J.-H. Kim, M. You, Q. Li, C. Fan, J.-M. Nam, Chem. Rev. 2019, 119, 12208.
[4] P. Jena, Q. Sun, Chem. Rev. 2018, 118, 5755.
[5] R. Ferrando, J. Phys. Condens. Matter 2014, 27, 013003.
[6] G. E. Johnson, R. Mitric, V. Bonačic-Koutecký, A. W. Castleman, Chem. Phys. Lett. 2009, 475, 1.
[7] L. Liu, A. Corma, Chem. Rev. 2018, 118, 4981.
[8] Y. Du, H. Sheng, D. Astruc, M. Zhu, Chem. Rev. 2020, 120, 526.
[9] L. E. VanGelder, A. M. Kosswattaarachchi, P. L. Forrestel, T. R. Cook, E. M. Matson, Chem. Sci. 2018, 9, 1692.
[10] L. E. VanGelder, E. Schreiber, E. M. Matson, J. Mater. Chem. A 2019, 7, 4893.
[11] L. E. VanGelder, H. D. Pratt, T. M. Anderson, E. M. Matson, Chem. Comm. 2019, 55, 12247.
[12] T.-H. Shin, Y. Choi, S. Kim, J. Cheon, Chem. Soc. Rev. 2015, 44, 4501.
[13] Y. Hu, S. Mignani, J.-P. Majoral, M. Shen, X. Shi, Chem. Soc. Rev. 2018, 47, 1874.
[14] D. Ni, W. Bu, E. B. Ehlerding, W. Cai, J. Shi, Chem. Soc. Rev. 2017, 46, 7438.
[15] J. T. O'Brien, J. S. Prell, M. F. Bush, E. R. Williams, J. Am. Chem. Soc. 2010, 132, 8248.
[16] Z. W. Wang, R. E. Palmer, Nano Lett. 2012, 12, 5510.
[17] D. J. Wales, J. P. K. Doye, J. Chem. Phys. 2003, 119, 12409.
[18] E. L. Altschuler, T. J. Williams, E. R. Ratner, R. Tipton, R. Stong, F. Dowla, F. Wooten, Phys. Rev. Lett. 1997, 78, 2681.
[19] J. Li, X. Li, H.-J. Zhai, L.-S. Wang, Science 2003, 299, 864.
[20] D. M. Bates, G. S. Tschumper, J. Phys. Chem. A 2009, 113, 3555.
[21] Y. Wang, V. Babin, J. M. Bowman, F. Paesani, J. Am. Chem. Soc. 2012, 134, 11116.
[22] J. Zhang, M. Dolg, J. Chem. Phys. 2014, 140, 044114.
[23] K. Liu, M. G. Brown, C. Carter, R. J. Saykally, J. K. Gregory, D. C. Clary, Nature 1996, 381, 501.
[24] C. Pérez, M. T. Muckle, D. P. Zaleski, N. A. Seifert, B. Temelso, G. C. Shields, Z. Kisiel, B. H. Pate, Science 2012, 336, 897.
[25] N. J. Rijs, P. González-Navarrete, M. Schlangen, H. Schwarz, J. Am. Chem. Soc. 2016, 138, 3125.
[26] G. Takasao, T. Wada, A. Thakur, P. Chammingkwan, M. Terano, T. Taniike, ACS Catal. 2019, 9, 2599.
[27] S. Heiles, R. L. Johnston, Int. J. Quantum Chem. 2013, 113, 2091.
[28] A. N. Alexandrova, A. I. Boldyrev, J. Chem. Theory Comput. 2005, 1, 566.
[29] S. Bulusu, S. Yoo, X. C. Zeng, J. Chem. Phys. 2005, 122, 164305.
[30] G. Barcaro, E. Aprà, A. Fortunelli, Chem. – Eur. J. 2007, 13, 6408.
[31] A. N. Alexandrova, J. Phys. Chem. A 2010, 114, 12591.
[32] L. Liu, J. Zhong, H. Vehkamäki, T. Kurtén, L. Du, X. Zhang, J. S. Francisco, X. C. Zeng, Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 24966.
[33] H. Zhai, P. Sautet, A. N. Alexandrova, ChemCatChem 2020, 12, 762.
[34] S. A. Akhade, A. Winkelman, V. Lebarbier Dagle, L. Kovarik, S. F. Yuk, M.-S. Lee, J. Zhang, A. B. Padmaperuma, R. A. Dagle, V.-A. Glezakou, Y. Wang,
R. Rousseau, J. Catal. 2020, 386, 30.
[35] J. P. K. Doye, D. J. Wales, R. S. Berry, J. Chem. Phys. 1995, 103, 4234.
[36] D. J. Wales, J. P. K. Doye, J. Phys. Chem. A 1997, 101, 5111.
[37] W. J. Pullan, J. Chem. Inf. Comput. Sci. 1997, 37, 1189.
[38] P. K. Doye, J.; J. Wales, D., . New J. Chem. 1998, 22, 733–744.
[39] R. P. White, J. A. Niesse, H. R. Mayne, J. Chem. Phys. 1998, 108, 2208.
[40] D. J. Wales, M. P. Hodges, Chem. Phys. Lett. 1998, 286, 65.
[41] D. Romero, C. Barrón, S. Gómez, Comput. Phys. Commun. 1999, 123, 87.
[42] J. P. K. Doye, D. J. Wales, Phys. Rev. B 1999, 59, 2292.
[43] V. Meleshko, Y. Morokov, V. Schweigert, Chem. Phys. Lett. 1999, 300, 118.
[44] M. P. Hodges, D. J. Wales, Chem. Phys. Lett. 2000, 324, 279.
[45] Calvo, F.; Boutin, A.; Labastie, P. In Structure of nitrogen molecular clusters (N2)n with 13 ≤ n ≤ 55, Eur. Phys. J. D, Berlin, Heidelberg, 1999//; Châtelain,
A.; Bonard, J. M., Eds. Springer Berlin Heidelberg: Berlin, Heidelberg, 1999; pp. 189–193.
[46] C. Roberts, R. L. Johnston, Phys. Chem. Chem. Phys. 2001, 3, 5024.
[47] F. Schulz, B. Hartke, ChemPhysChem 2002, 3, 98.
[48] J. P. K. Doye, Phys. Rev. B 2003, 68, 195418.
[49] G. Rossi, A. Rapallo, C. Mottet, A. Fortunelli, F. Baletto, R. Ferrando, Phys. Rev. Lett. 2004, 93, 105503.
[50] S. Yoo, J. Zhao, J. Wang, X. C. Zeng, J. Am. Chem. Soc. 2004, 126, 13845.
[51] T. James, D. J. Wales, J. Hernández-Rojas, Chem. Phys. Lett. 2005, 415, 302.
[52] B. S. González, J. Hernández-Rojas, D. J. Wales, Chem. Phys. Lett. 2005, 412, 23.
[53] S. T. Bromley, E. Flikkema, Phys. Rev. Lett. 2005, 95, 185505.
ZHANG AND GLEZAKOU 15 of 18

[54] A. Sebetci, Z. B. Güvenç, Modell. Simul. Mater. Sci. Eng. 2005, 13, 683.
[55] A. Costales, M. A. Blanco, E. Francisco, R. Pandey, A. Martín Pendás, J. Phys. Chem. B 2005, 109, 24352.
[56] J. Hernández-Rojas, J. Bretón, J. M. Gomez Llorente, D. J. Wales, J. Phys. Chem. B 2006, 110, 13357.
[57] E. Oger, N. R. M. Crawford, R. Kelting, P. Weis, M. M. Kappes, R. Ahlrichs, Angew. Chem., Int. Ed. 2007, 46, 8503.
[58] B. S. González, J. Hernández-Rojas, J. Bretón, J. M. Gomez Llorente, J. Phys. Chem. C 2008, 112, 16497.
[59] B. Assadollahzadeh, P. R. Bunker, P. Schwerdtfeger, Chem. Phys. Lett. 2008, 451, 262.
[60] B. Assadollahzadeh, C. Thierfelder, P. Schwerdtfeger, Phys. Rev. B 2008, 78, 245423.
[61] G. Santambrogio, E. Janssens, S. Li, T. Siebert, G. Meijer, K. R. Asmis, J. Döbler, M. Sierka, J. Sauer, J. Am. Chem. Soc. 2008, 130, 15143.
[62] B. Assadollahzadeh, P. Schwerdtfeger, J. Chem. Phys. 2009, 131, 064306.
[63] E. Oger, R. Kelting, P. Weis, A. Lechtken, D. Schooss, N. R. M. Crawford, R. Ahlrichs, M. M. Kappes, J. Chem. Phys. 2009, 130, 124305.
[64] H. Takeuchi, J. Comput. Chem. 2010, 31, 1699.
[65] D.-E. Jiang, M. Walter, S. Dai, Chem. – Eur. J. 2010, 16, 4999.
[66] H. Xiang, S.-H. Wei, X. Gong, J. Am. Chem. Soc. 2010, 132, 7355.
[67] P. Karamanis, R. Marchal, P. Carbonniére, C. Pouchan, J. Chem. Phys. 2011, 135, 044511.
[68] H. Do, N. A. Besley, J. Chem. Phys. 2012, 137, 134106.
[69] L. B. Vilhelmsen, B. Hammer, Phys. Rev. Lett. 2012, 108, 126101.
[70] Z. A. Piazza, W.-L. Li, C. Romanescu, A. P. Sergeeva, L.-S. Wang, A. I. Boldyrev, J. Chem. Phys. 2012, 136, 104310.
[71] J. L. Llanio-Trujillo, J. M. C. Marques, F. B. Pereira, Comput. Theor. Chem. 2013, 1021, 124.
[72] M. Chen, D. A. Dixon, Nanoclusters. J. Chem. Theory Comput. 2013, 9, 3189.
[73] Y.-R. Liu, H. Wen, T. Huang, X.-X. Lin, Y.-B. Gai, C.-J. Hu, W.-J. Zhang, W. Huang, J. Phys. Chem. A 2014, 118, 508.
[74] F. R. Negreiros, S. Fabris, J. Phys. Chem. C 2014, 118, 21014.
[75] C. Priest, Q. Tang, D.-E. Jiang, J. Phys. Chem. A 2015, 119, 8892.
[76] J. B. A. Davis, A. Shayeghi, S. L. Horswell, R. L. Johnston, Nanoscale 2015, 7, 14032.
[77] L. C. Smeeton, J. D. Farrell, M. T. Oakley, D. J. Wales, R. L. Johnston, J. Chem. Theory Comput. 2015, 11, 2377.
[78] Y. Feng, L. Cheng, RSC Adv. 2015, 5, 62543.
[79] N. X. Truong, M. Savoca, D. J. Harding, A. Fielicke, O. Dopfer, Phys. Chem. Chem. Phys. 2015, 17, 18961.
[80] L. B. Vilhelmsen, B. Hammer, ACS Catal. 2014, 4, 1626.
[81] L. P. Ding, F. H. Zhang, Y. S. Zhu, C. Lu, X. Y. Kuang, J. Lv, P. Shao, Sci. Rep. 2015, 5, 15951.
[82] Y. Jin, G. Maroulis, X. Kuang, L. Ding, C. Lu, J. Wang, J. Lv, C. Zhang, M. Ju, Phys. Chem. Chem. Phys. 2015, 17, 13590.
[83] X. Xing, A. Hermann, X. Kuang, M. Ju, C. Lu, Y. Jin, X. Xia, G. Maroulis, Sci. Rep. 2016, 6, 19656.
[84] G.-F. Wei, Z.-P. Liu, J. Chem. Theory Comput. 2016, 12, 4698.
[85] M. Aslan, J. B. A. Davis, R. L. Johnston, Phys. Chem. Chem. Phys. 2016, 18, 6676.
[86] J. C. Hey, L. C. Smeeton, M. T. Oakley, R. L. Johnston, J. Phys. Chem. A 2016, 120, 4008.
[87] W. G. Sun, J. J. Wang, C. Lu, X. X. Xia, X. Y. Kuang, A. Hermann, Inorg. Chem. 2017, 56, 1241.
[88] C. Siouani, S. Mahtout, S. Safer, F. Rabilloud, J. Phys. Chem. A 2017, 121, 3540.
[89] I. Demiroglu, K. Yao, H. A. Hussein, R. L. Johnston, J. Phys. Chem. C 2017, 121, 10773.
[90] A. Malloum, J. J. Fifen, J. Conradie, J. Chem. Phys. 2018, 149, 024304.
[91] Y. Liu, J. Yang, L. Cheng, Inorg. Chem. 2018, 57, 12934.
[92] S. Pande, X. Gong, L.-S. Wang, X. C. Zeng, J. Phys. Chem. Lett. 2019, 10, 1820.
[93] B. L. Chen, W. G. Sun, X. Y. Kuang, C. Lu, X. X. Xia, H. X. Shi, G. Maroulis, Inorg. Chem. 2018, 57, 343.
[94] M. Lippe, U. Szczepaniak, G.-L. Hou, S. Chakrabarty, J. J. Ferreiro, E. Chasovskikh, R. Signorell, J. Phys. Chem. A 2019, 123, 2426.
[95] X. Zhang, K. Jiang, Z. Liu, X. Yao, X. Liu, S. Zeng, K. Dong, S. Zhang, Ind. Eng. Chem. Res. 2019, 58, 1443.
[96] L. Ahonen, C. Li, J. Kubečka, S. Iyer, H. Vehkamäki, T. Petäjä, M. Kulmala, C. J. Hogan Jr., J. Phys. Chem. Lett. 2019, 10, 1935.
[97] A. S. Bazhenov, K. Honkala, J. Phys. Chem. C 2019, 123, 7209.
[98] R. L. Johnston, Dalt. Trans. 2003, 22, 4193.
[99] M. Jäger, R. Schäfer, R. L. Johnston, Adv. Phys.: X 2018, 3, 1516514.
[100] D. J. Wales, H. A. Scheraga, Science 1999, 285, 1368.
[101] J. P. K. Doye, D. J. Wales, Phys. Rev. Lett. 1998, 80, 1357.
[102] C. Shang, Z.-P. Liu, J. Chem. Theory Comput. 2013, 9, 1838.
[103] A. Laio, M. Parrinello, Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 12562.
[104] J. P. K. Doye, M. A. Miller, D. J. Wales, J. Chem. Phys. 1999, 111, 8417.
[105] J. P. K. Doye, D. J. Wales, F. H. M. Zetterling, M. Dzugutov, J. Chem. Phys. 2003, 118, 2792.
[106] J. Zhang, M. Dolg, Phys. Chem. Chem. Phys. 2016, 18, 3003.
[107] C. Bannwarth, S. Ehlert, S. Grimme, J. Chem. Theory Comput. 2019, 15, 1652.
[108] P. J. Stephens, F. Devlin, C. Chabalowski, M. J. Frisch, J. Phys. Chem. 1994, 98, 11623.
[109] C. Riplinger, B. Sandhoefer, A. Hansen, F. Neese, J. Chem. Phys. 2013, 139, 134101.
[110] B. R. L. Galv~ao, L. P. Viegas, J. Phys. Chem. A 2019, 123, 10454.
[111] R. H. Leary, J. Global Optim. 2000, 18, 367.
[112] H. G. Kim, S. K. Choi, H. M. Lee, J. Chem. Phys. 2008, 128, 144702.
[113] L. Zhan, J. Z. Y. Chen, W.-K. Liu, S. K. Lai, J. Chem. Phys. 2005, 122, 244707.
[114] Y. Zhao, X. Chen, J. Li, Nano Res. 2017, 10, 3407.
[115] L. T. Wille, Chem. Phys. Lett. 1987, 133, 405.
[116] Z. Li, H. A. Scheraga, J. Mol. Struct.: THEOCHEM 1988, 179, 333.
[117] Y. Xiang, L. Cheng, W. Cai, X. Shao, J. Phys. Chem. A 2004, 108, 9516.
[118] K. Yu, X. Wang, L. Chen, L. Wang, J. Chem. Phys. 2019, 151, 214105.
16 of 18 ZHANG AND GLEZAKOU

[119] J. Zhang, M. Dolg, Phys. Chem. Chem. Phys. 2015, 17, 24173.
[120] X. Chen, Y.-F. Zhao, Y.-Y. Zhang, J. Li, J. Comput. Chem. 2019, 40, 1105.
[121] M. I. Aroyo, J. Perez-Mato, D. Orobengoa, E. Tasci, G. De La Flor, A. Kirov, Bulg. Chem. Commun. 2011, 43, 183.
[122] W.-L. Li, T. Jian, X. Chen, T.-T. Chen, G. V. Lopez, J. Li, L.-S. Wang, Angew. Chem., Int. Ed. 2016, 55, 7358.
[123] B. Kiran, S. Bulusu, H.-J. Zhai, S. Yoo, X. C. Zeng, L.-S. Wang, Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 961.
[124] M. Rapacioli, F. Calvo, F. Spiegelman, C. Joblin, D. J. Wales, J. Phys. Chem. A 2005, 109, 2487.
[125] C. Shang, Z.-P. Liu, J. Chem. Theory Comput. 2012, 8, 2215.
[126] N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, E. Teller, J. Chem. Phys. 1953, 21, 1087.
[127] M. Tuckerman, Statistical mechanics: theory and molecular simulation, Oxford university press, Oxford 2010.
[128] J. H. Holland, Adaptation in natural and artificial systems: an introductory analysis with applications to biology, control, and artificial intelligence, MIT
press, Cambridge 1992.
[129] B. Hartke, J. Phys. Chem. 1993, 97, 9973.
[130] D. M. Deaven, K. M. Ho, Phys. Rev. Lett. 1995, 75, 288.
[131] B. Hartke, Chem. Phys. Lett. 1995, 240, 560.
[132] D. M. Daven, N. Tit, J. R. Morris, K. M. Ho, Chem. Phys. Lett. 1996, 256, 195.
[133] R. P. F. Kanters, K. J. Donald, J. Chem. Theory Comput. 2014, 10, 5729.
[134] A. Shayeghi, D. Götz, J. B. A. Davis, R. Schäfer, R. L. Johnston, Phys. Chem. Chem. Phys. 2015, 17, 2104.
[135] W. A. Rabanal-León, W. Tiznado, E. Osorio, F. Ferraro, RSC Adv. 2018, 8, 145.
[136] M. Jäger, R. Schäfer, R. L. Johnston, Nanoscale 2019, 11, 9042.
[137] O. Yañez, R. Báez-Grez, D. Inostroza, W. A. Rabanal-León, R. Pino-Rios, J. Garza, W. Tiznado, J. Chem. Theory Comput. 2019, 15, 1463.
[138] T. Lazauskas, A. A. Sokol, S. M. Woodley, Nanoscale 2017, 9, 3850.
[139] J. Rogan, A. Varas, J. A. Valdivia, M. Kiwi, J. Comput. Chem. 2013, 34, 2548.
[140] J. Zhang, V.-A. Glezakou, R. Rousseau, M.-T. Nguyen, J. Chem. Theory Comput. 2020, 16, 3947.
[141] Z. Chen, X. Jiang, J. Li, S. Li, L. Wang, J. Comput. Chem. 2013, 34, 1046.
[142] Z. Chen, W. Jia, X. Jiang, S.-S. Li, L.-W. Wang, Comput. Phys. Commun. 2017, 219, 35.
[143] Y.-H. Yang, X.-B. Xu, S.-B. He, J.-B. Wang, Y.-H. Wen, Comput. Mater. Sci. 2018, 149, 416.
[144] T.-E. Fan, G.-F. Shao, Q.-S. Ji, J.-W. Zheng, T.-D. Liu, Y.-H. Wen, Comput. Phys. Commun. 2016, 208, 64.
[145] H. Zhai, M.-A. Ha, A. N. Alexandrova, J. Chem. Theory Comput. 2015, 11, 2385.
[146] M. A. Addicoat, G. F. Metha, J. Comput. Chem. 2009, 30, 57.
[147] Bera, P. P.; Schleyer, P. v. R.; Schaefer, H. F., . Int. J. Quantum Chem. 2007, 107, 2220–2223.
[148] M. Saunders, J. Am. Chem. Soc. 1987, 109, 3150.
[149] S. T. Call, D. Y. Zubarev, A. I. Boldyrev, J. Comput. Chem. 2007, 28, 1177.
[150] G.-F. Shao, J.-B. Yang, J.-F. Zhang, T.-D. Liu, Y.-H. Wen, Phys. Lett. A 2019, 383, 3123.
[151] J. Lv, Y. Wang, L. Zhu, Y. Ma, J. Chem. Phys. 2012, 137, 084104.
[152] B. Yan, Z. Zhao, Y. Zhou, W. Yuan, J. Li, J. Wu, D. Cheng, Comput. Phys. Commun. 2017, 219, 79.
[153] Y. Wang, J. Lv, L. Zhu, Y. Ma, Comput. Phys. Commun. 2012, 183, 2063.
[154] C. Roberts, R. L. Johnston, N. T. Wilson, Theor. Chem. Acc. 2000, 104, 123.
[155] Y. Zeiri, Phys. Rev. E 1995, 51, R2769.
[156] X. F. Yan, D. Z. Chen, S. X. Hu, Comput. Chem. Eng. 2003, 27, 1393.
[157] Karaboga, D., An Idea Based on Honey Bee Swarm for Numerical Optimization. Technical Report TR06, Erciyes University: 2005.
[158] T. D. Seeley, The wisdom of the hive: the social physiology of honey bee colonies, Harvard University Press, Cambridge 2009.
[159] H.-Y. Fan, J. Lampinen, J. Global Optim. 2003, 27, 105.
[160] W. Gao, S. Liu, Inform. Process. Lett. 2011, 111, 871.
[161] E. Bonabeau, M. Dorigo, D. d. R. D. F. Marco, G. Theraulaz, G. Théraulaz, Swarm intelligence: from natural to artificial systems, Oxford university press,
Oxford 1999.
[162] Y. Xiao, D. E. Williams, Chem. Phys. Lett. 1993, 215, 17.
[163] http://www-wales.ch.cam.ac.uk/GMIN/.
[164] V.-M. Kerminen, X. Chen, V. Vakkari, T. Petäjä, M. Kulmala, F. Bianchi, Environ. Res. Lett. 2018, 13, 103003.
[165] H.-B. Xie, J. Elm, R. Halonen, N. Myllys, T. Kurtén, M. Kulmala, H. Vehkamäki, Environ. Sci. Technol. 2017, 51, 8422.
[166] C. Dameto de España, A. Wonaschütz, G. Steiner, B. Rosati, A. Demattio, H. Schuh, R. Hitzenberger, Atmos. Environ. 2017, 164, 289.
[167] M. Kumar, H. Li, X. Zhang, X. C. Zeng, J. S. Francisco, J. Am. Chem. Soc. 2018, 140, 6456.
[168] Change, I. C., The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate
Change. 2013 2013, 33–118.
[169] C. A. Pope, D. W. Dockery, J. Air Waste Manage. Assoc. 2006, 56, 709.
[170] I. K. Ortega, O. Kupiainen, T. Kurtén, T. Olenius, O. Wilkman, M. J. McGrath, V. Loukonen, H. Vehkamäki, Atmos. Chem. Phys. 2012, 12, 225.
[171] M. J. McGrath, T. Olenius, I. K. Ortega, V. Loukonen, P. Paasonen, T. Kurtén, M. Kulmala, H. Vehkamäki, Atmos. Chem. Phys. 2012, 12, 2345.
[172] G.-L. Hou, J. Zhang, M. Valiev, X.-B. Wang, Phys. Chem. Chem. Phys. 2017, 19, 10676.
[173] J. Ling, X. Ding, Z. Li, J. Yang, J. Phys. Chem. A 2017, 121, 661.
[174] H. Zhang, O. Kupiainen-Määttä, X. Zhang, V. Molinero, Y. Zhang, Z. Li, J. Chem. Phys. 2017, 146, 184308.
[175] L. Liu, O. Kupiainen-Määttä, H. Zhang, H. Li, J. Zhong, T. Kurtén, H. Vehkamäki, S. Zhang, Y. Zhang, M. Ge, X. Zhang, Z. Li, J. Chem. Phys. 2018, 148,
214303.
[176] H. Li, J. Zhong, H. Vehkamäki, T. Kurtén, W. Wang, M. Ge, S. Zhang, Z. Li, X. Zhang, J. S. Francisco, X. C. Zeng, J. Am. Chem. Soc. 2018, 140, 11020.
[177] H. Zhang, W. Wang, S. Pi, L. Liu, H. Li, Y. Chen, Y. Zhang, X. Zhang, Z. Li, Chemosphere 2018, 212, 504.
[178] H. Zhang, H. Li, L. Liu, Y. Zhang, X. Zhang, Z. Li, Chemosphere 2018, 203, 26.
[179] X. Ma, Y. Sun, Z. Huang, Q. Zhang, W. Wang, Chemosphere 2019, 214, 781.
ZHANG AND GLEZAKOU 17 of 18

[180] H. Li, X. Zhang, J. Zhong, L. Liu, H. Zhang, F. Chen, Z. Li, Q. Li, M. Ge, Atmos. Environ. 2018, 189, 244.
[181] H. Li, O. Kupiainen-Määttä, H. Zhang, X. Zhang, M. Ge, Atmos. Environ. 2017, 166, 479.
[182] L. Liu, H. Li, H. Zhang, J. Zhong, Y. Bai, M. Ge, Z. Li, Y. Chen, X. Zhang, Phys. Chem. Chem. Phys. 2018, 20, 17406.
[183] N. Myllys, S. Chee, T. Olenius, M. Lawler, J. Smith, J. Phys. Chem. A 2019, 123, 2420.
[184] D. Li, D. Chen, F. Liu, W. Wang, J. Environ. Sci. 2020, 89, 125.
[185] H. Wang, X. Zhao, C. Zuo, X. Ma, F. Xu, Y. Sun, Q. Zhang, RSC Adv. 2019, 9, 36171.
[186] R. R. Valiev, G. Hasan, V.-T. Salo, J. Kubečka, T. Kurten, J. Phys. Chem. A 2019, 123, 6596.
[187] N. Myllys, J. Kubečka, V. Besel, D. Alfaouri, T. Olenius, J. N. Smith, M. Passananti, Atmos. Chem. Phys. 2019, 19, 9753.
[188] J. Kubečka, V. Besel, T. Kurtén, N. Myllys, H. Vehkamäki, J. Phys. Chem. A 2019, 123, 6022.
[189] F. Keshavarz, A. Shcherbacheva, J. Kubečka, H. Vehkamäki, T. Kurtén, J. Phys. Chem. A 2019, 123, 9008.
[190] S. Chee, N. Myllys, K. C. Barsanti, B. M. Wong, J. N. Smith, J. Phys. Chem. A 2019, 123, 5640.
[191] D. Chen, D. Li, C. Wang, Y. Luo, F. Liu, W. Wang, Chemosphere 2020, 244, 125538.
[192] D. Chen, D. Li, C. Wang, F. Liu, W. Wang, Atmos. Environ. 2020, 222, 117161.
[193] D. Chen, W. Wang, D. Li, W. Wang, RSC Adv. 2020, 10, 5173.
[194] S. Ni, F.-Y. Bai, X.-M. Pan, Phys. Chem. Chem. Phys. 2020, 22, 8109.
[195] A. Ning, H. Zhang, X. Zhang, Z. Li, Y. Zhang, Y. Xu, M. Ge, Atmos. Environ. 2020, 227, 117378.
[196] H. Rong, J. Liu, Y. Zhang, L. Du, X. Zhang, Z. Li, Chemosphere 2020, 253, 126743.
[197] H. Rong, L. Liu, J. Liu, X. Zhang, J. Phys. Chem. A 2020, 124, 3261.
[198] V. Besel, J. Kubečka, T. Kurtén, H. Vehkamäki, J. Phys. Chem. A 2020, 124, 5931.
[199] A. Leonardi, H. M. Ricker, A. G. Gale, B. T. Ball, T. T. Odbadrakh, G. C. Shields, J. G. Navea, Int. J. Quantum Chem. 2020, 120, e26350.
[200] J. Elm, J. Kubečka, V. Besel, M. J. Jääskeläinen, R. Halonen, T. Kurtén, H. Vehkamäki, J. Aerosol Sci. 2020, 149, 105621.
[201] M. Kulmala, T. Petäjä, M. Ehn, J. Thornton, M. Sipilä, D. R. Worsnop, V. M. Kerminen, Ann. Rev. Phys. Chem. 2014, 65, 21.
[202] N. Tarrat, M. Rapacioli, J. Cuny, J. Morillo, J.-L. Heully, F. Spiegelman, Comput. Theor. Chem. 2017, 1107, 102.
[203] M. Van den Bossche, J. Phys. Chem. A 2019, 123, 3038.
[204] Häkkinen, H.; Moseler, M.; Kostko, O.; Morgner, N.; Hoffmann, M. A.; v. Issendorff, B., . Phys. Rev. Lett. 2004, 93, 093401.
[205] S. Pande, W. Huang, N. Shao, L.-M. Wang, N. Khetrapal, W.-N. Mei, T. Jian, L.-S. Wang, X. C. Zeng, ACS Nano 2016, 10, 10013.
[206] B. Schaefer, R. Pal, N. S. Khetrapal, M. Amsler, A. Sadeghi, V. Blum, X. C. Zeng, S. Goedecker, L.-S. Wang, ACS Nano 2014, 8, 7413.
[207] N. S. Khetrapal, S. S. Bulusu, X. C. Zeng, J. Phys. Chem. A 2017, 121, 2466.
[208] Y. Chen, Y. Liu, S. Li, J. Yang, J. Cluster Sci. 2019, 30, 789.
[209] S. He, J. Yang, J. Cluster Sci. 2017, 28, 2309.
[210] Y. Zhang, J. Yang, L. Cheng, J. Cluster Sci. 2018, 29, 301.
[211] L. Hou, J. Yang, Y. Liu, J. Mol. Model. 2017, 23, 117.
[212] X. Huang, J. Yang, J. Mol. Model. 2017, 24, 29.
[213] Y. Liu, J. Yang, S. Li, L. Cheng, RSC Adv. 2019, 9, 2731.
[214] Y. Feng, J. Yang, Y. Liu, Theor. Chem. Acc. 2016, 135, 258.
[215] S. He, J. Yang, Theor. Chem. Acc. 2017, 136, 93.
[216] J. Yang, Y. Feng, X. Xie, H. Wu, Y. Liu, Theor. Chem. Acc. 2016, 135, 204.
[217] Y. Chen, J. Yang, C. Dong, Comput. Theor. Chem. 2019, 1170, 112635.
[218] C. Dong, L. Han, J. Yang, L. Cheng, Int. J. Mol. Sci. 2019, 20, 2933.
[219] C. Dong, J. Yang, J. Lu, J. Mol. Model. 2020, 26, 85.
[220] Y. Gu, J. Yang, L. Cheng, Int. J. Quantum Chem. 2020, 120, e26087.
[221] N. S. Khetrapal, T. Jian, G. V. Lopez, S. Pande, L.-S. Wang, X. C. Zeng, J. Phys. Chem. C 2017, 121, 18234.
[222] N. S. Khetrapal, T. Jian, R. Pal, G. V. Lopez, S. Pande, L.-S. Wang, X. C. Zeng, Nanoscale 2016, 8, 9805.
[223] S. Pande, T. Jian, N. S. Khetrapal, L.-S. Wang, X. C. Zeng, J. Phys. Chem. C 2018, 122, 6947.
[224] X.-M. Luo, T. Jian, L.-J. Cheng, W.-L. Li, Q. Chen, R. Li, H.-J. Zhai, S.-D. Li, A. I. Boldyrev, J. Li, L.-S. Wang, Chem. Phys. Lett. 2017, 683, 336.
[225] L.-S. Wang, Int. Rev. Phys. Chem. 2016, 35, 69.
[226] Q. Chen, W.-J. Tian, L.-Y. Feng, H.-G. Lu, Y.-W. Mu, H.-J. Zhai, S.-D. Li, L.-S. Wang, Nanoscale 2017, 9, 4550.
[227] W.-L. Li, Q. Chen, W.-J. Tian, H. Bai, Y.-F. Zhao, H.-S. Hu, J. Li, H.-J. Zhai, S.-D. Li, L.-S. Wang, J. Am. Chem. Soc. 2014, 136, 12257.
[228] H. Bai, T.-T. Chen, Q. Chen, X.-Y. Zhao, Y.-Y. Zhang, W.-J. Chen, W.-L. Li, L. F. Cheung, B. Bai, J. Cavanagh, W. Huang, S.-D. Li, J. Li, L.-S. Wang,
Nanoscale 2019, 11, 23286.
[229] H.-J. Zhai, Y.-F. Zhao, W.-L. Li, Q. Chen, H. Bai, H.-S. Hu, Z. A. Piazza, W.-J. Tian, H.-G. Lu, Y.-B. Wu, Y.-W. Mu, G.-F. Wei, Z.-P. Liu, J. Li, S.-D. Li, L.-
S. Wang, Nat. Chem. 2014, 6, 727.
[230] T.-T. Chen, W.-L. Li, H. Bai, W.-J. Chen, X.-R. Dong, J. Li, L.-S. Wang, J. Phys. Chem. A 2019, 123, 5317.
[231] W.-L. Li, T. Jian, X. Chen, H.-R. Li, T.-T. Chen, X.-M. Luo, S.-D. Li, J. Li, L.-S. Wang, Chem. Comm. 2017, 53, 1587.
[232] T. Jian, W.-L. Li, I. A. Popov, G. V. Lopez, X. Chen, A. I. Boldyrev, J. Li, L.-S. Wang, J. Chem. Phys. 2016, 144, 154310.
[233] P. J. Robinson, X. Zhang, T. McQueen, K. H. Bowen, A. N. Alexandrova, J. Phys. Chem. A 2017, 121, 1849.
[234] T.-T. Chen, W.-L. Li, T. Jian, X. Chen, J. Li, L.-S. Wang, Angew. Chem., Int. Ed. 2017, 56, 6916.
[235] W.-L. Li, T.-T. Chen, D.-H. Xing, X. Chen, J. Li, L.-S. Wang, Proc. Natl. Acad. Sci. U. S. A. 2018, 115, E6972.
[236] J. J. Torres-Vega, A. Vásquez-Espinal, M. J. Beltran, L. Ruiz, R. Islas, W. Tiznado, Phys. Chem. Chem. Phys. 2015, 17, 19602.
[237] N. Perez-Peralta, M. Contreras, W. Tiznado, J. Stewart, K. J. Donald, G. Merino, Phys. Chem. Chem. Phys. 2011, 13, 12975.
[238] W. Tiznado, N. Perez-Peralta, R. Islas, A. Toro-Labbe, J. M. Ugalde, G. Merino, J. Am. Chem. Soc. 2009, 131, 9426.
[239] M. Contreras, E. Osorio, F. Ferraro, G. Puga, K. J. Donald, J. G. Harrison, G. Merino, W. Tiznado, Chem. – Eur. J. 2013, 19, 2305.
[240] S. Jin, B. Chen, X. Kuang, C. Lu, W. Sun, X. Xia, G. L. Gutsev, J. Phys. Chem. C 2019, 123, 6276.
[241] W. Sun, X. Xia, C. Lu, X. Kuang, A. Hermann, Phys. Chem. Chem. Phys. 2018, 20, 23740.
[242] Y. Tian, D. Wei, Y. Jin, J. Barroso, C. Lu, G. Merino, Phys. Chem. Chem. Phys. 2019, 21, 6935.
18 of 18 ZHANG AND GLEZAKOU

[243] C. W. Glass, A. R. Oganov, N. Hansen, Comput. Phys. Commun. 2006, 175, 713.
[244] L. Shao, K. Wan, H. Wang, Y. Cui, C. Zhao, J. Lu, X. Li, L. Chen, X. Cui, X. Wang, X. Deng, X. Shi, Y. Wu, Biomater. Sci. 2019, 7, 3016.
[245] J. Wu, L.-W. Wang, J. Mater. Chem. A 2018, 6, 2984.
[246] X. Ma, S. Liu, S. Huang, Int. J. Hydrogen Energ. 2017, 42, 24797.
[247] Y. Xu, N. Wang, X. Guo, S. Huang, Int. J. Hydrogen Energ. 2019, 44, 31029.
[248] G. Chen, C. Xu, X. Huang, J. Ye, L. Gu, G. Li, Z. Tang, B. Wu, H. Yang, Z. Zhao, Z. Zhou, G. Fu, N. Zheng, Nat. Mater. 2016, 15, 564.
[249] J. Laskin, G. E. Johnson, J. Warneke, V. Prabhakaran, Angew. Chem., Int. Ed. 2018, 57, 16270.
[250] X. Roy, C.-H. Lee, A. C. Crowther, C. L. Schenck, T. Besara, R. A. Lalancette, T. Siegrist, P. W. Stephens, L. E. Brus, P. Kim, M. L. Steigerwald, C.
Nuckolls, Science 2013, 341, 157.
[251] V. S. Bryantsev, M. S. Diallo, W. A. Goddard Iii, J. Phys. Chem. B 2008, 112, 9709.
[252] A. Malloum, J. J. Fifen, J. Conradie, J. Chem. Phys. 2018, 149, 244301.
[253] A. Malloum, J. J. Fifen, J. Conradie, J. Comput. Chem. 2020, 41, 21.
[254] A. Malloum, J. J. Fifen, J. Conradie, Int. J. Quantum Chem. 2020, 120, e26221.
[255] A. Malloum, J. J. Fifen, J. Conradie, Int. J. Quantum Chem. 2020, 120, e26234.
[256] A. Malloum, J. J. Fifen, J. Conradie, Phys. Chem. Chem. Phys. 2020, 22, 13201.
[257] Z. Wu, K. Huang, X. Kuang, RSC Adv. 2016, 6, 66007.
[258] X. Cao, N. Heinz, J. Zhang, M. Dolg, Theoretical studies. Phys. Chem. Chem. Phys. 2017, 19, 20160.
[259] J. Zhang, E. T. Baxter, M.-T. Nguyen, V. Prabhakaran, R. Rousseau, G. E. Johnson, V.-A. Glezakou, J. Phys. Chem. Lett. 2020, 11, 6844.
[260] J. Li, S. Zhou, J. Zhang, M. Schlangen, D. Usharani, S. Shaik, H. Schwarz, J. Am. Chem. Soc. 2016, 138, 11368.
[261] S. K. Ignatov, A. I. Okhapkin, O. B. Gadzhiev, A. G. Razuvaev, S. Kunz, M. Bäumer, A DFT Study. J. Phys. Chem. C 2016, 120, 18570.
[262] V. T. Tran, Q. T. Tran, J. Phys. Chem. A 2019, 123, 9223.
[263] T. H. Tran, Q. T. Tran, V. T. Tran, Comput. Theor. Chem. 2020, 1173, 112701.
[264] Y. Basdogan, J. A. Keith, Chem. Sci. 2018, 9, 5341.
[265] Y. Basdogan, M. C. Groenenboom, E. Henderson, S. De, S. B. Rempe, J. A. Keith, J. Chem. Theory Comput. 2020, 16, 633.
[266] A. M. Maldonado, Y. Basdogan, J. T. Berryman, S. B. Rempe, J. A. Keith, J. Chem. Phys. 2020, 152, 130902.
[267] A. F. Roesel, M. Ugandi, N. T. T. Huyen, M. Májek, T. Broese, M. Roemelt, R. Francke, J. Org. Chem. 2020, 85, 8029.
[268] Y.-G. Wang, Y. Yoon, V.-A. Glezakou, J. Li, R. Rousseau, J. Am. Chem. Soc. 2013, 135, 10673.
[269] N. O. B. Lüttschwager, T. N. Wassermann, R. A. Mata, M. A. Suhm, Angew. Chem., Int. Ed. 2013, 52, 463.
[270] P. C. D. Hawkins, J. Chem. Inf. Model. 2017, 57, 1747.
[271] K. Konishi, M. Iwasaki, Y. Shichibu, Acc. Chem. Res. 2018, 51, 3125.
[272] P. D. Jadzinsky, G. Calero, C. J. Ackerson, D. A. Bushnell, R. D. Kornberg, Science 2007, 318, 430.
[273] Y. Pei, Y. Gao, N. Shao, X. C. Zeng, J. Am. Chem. Soc. 2009, 131, 13619.
[274] J. Behler, J. Chem. Phys. 2016, 145, 170901.
[275] R. Ouyang, Y. Xie, D.-E. Jiang, Nanoscale 2015, 7, 14817.
[276] P. Tuo, X. B. Ye, B. C. Pan, J. Chem. Phys. 2020, 152, 114105.
[277] J. Behler, M. Parrinello, Phys. Rev. Lett. 2007, 98, 146401.
[278] Zhang, L.; Han, J.; Wang, H.; Car, R.; E, W., . Phys. Rev. Lett. 2018, 120, 143001.
[279] Y. Zhang, C. Hu, B. Jiang, J. Phys. Chem. Lett. 2019, 10, 4962.
[280] T. L. Jacobsen, M. S. Jørgensen, B. Hammer, Phys. Rev. Lett. 2018, 120, 026102.
[281] E. L. Kolsbjerg, A. A. Peterson, B. Hammer, Phys. Rev. B 2018, 97, 195424.
[282] M. K. Bisbo, B. Hammer, Phys. Rev. Lett. 2020, 124, 086102.
[283] M. S. Jørgensen, H. L. Mortensen, S. A. Meldgaard, E. L. Kolsbjerg, T. L. Jacobsen, K. H. Sørensen, B. Hammer, J. Chem. Phys. 2019, 151, 054111.
[284] S. A. Meldgaard, H. L. Mortensen, M. S. Jørgensen, B. Hammer, J. Phys. Condens. Matter 2020, 32, 404005.

AUTHOR BIOGRAPHY

Jun Zhang is a postdoctoral researcher with Drs V.-A. Glezakou and R. Rousseau at Pacific Northwest National Laboratory (PNNL). He earned
his bachelor's and master's degrees from Nankai University, then obtained his Ph.D. in Theoretical Chemistry at University of Cologne under
the supervision of Prof. Michael Dolg. He joined PNNL after postdoctoral work at University of Illinois at Urbana Champaign with Prof. So
Hirata. He is the recipient of 2014 GCCCD Young Researchers Award, 2017 MolSSI Fellowship, and Wiley Computers in Chemistry Out-
standing Postdoc Award for Spring 2020. His research interests include structure predictions, ab initio many-body methods, machine learning,
and their applications in chemistry.

How to cite this article: Zhang J, Glezakou V-A. Global optimization of chemical cluster structures: Methods, applications, and challenges.
Int J Quantum Chem. 2021;121:e26553. https://doi.org/10.1002/qua.26553

You might also like