Keywords

1 Introduction

Organic molecular crystals often exhibit a variety of different packing motifs, or polymorphs. These different crystal packing motifs can have diverse physical properties, making crystal structure critically important in a wide range of fields. Polymorphism plays a major role in the pharmaceutical industry, for example, where a substantial fraction of drugs including aspirin, acetaminophen, Lipitor, and Zantac have known polymorphs. Polymorphs of a given pharmaceutical can have drastically different solubilities and bioavailabilities, making the understanding of polymorphism critical for the drug industry.

Pharmaceutical polymorphism has led to several major drug recalls or withdrawals in recent years. For instance, the HIV drug ritonavir was temporarily removed from the market in 1998 when a new, insoluble polymorph appeared in production facilities, leading to shortages of this desperately needed medicine and costing its maker hundreds of millions of dollars in lost sales [1, 2]. Polymorphism is also believed to be behind multiple recalls of the anti-seizure drug carbamazepine, which exhibits several low-solubility polymorphs [3]. In 2008, Neupro brand skin patches for the Parkinson’s disease drug rotigotine were withdrawn from the market when a less-effective crystal form appeared visibly on the patches as dendritic structures [4]. Pharmaceutical polymorphism has also been the subject of many legal battles arising from the fact that unique crystal forms are patentable. Major examples include the ulcer/heartburn medication Zantac and the antibiotic cefadroxil [5].

Crystal packing is also important for foods such as chocolate. Solid cocoa butter form V is desired to achieve chocolate with a shiny appearance, a 34°C melting point that causes it to melt pleasingly on the tongue, and other favorable characteristics. However, form VI cocoa butter, which leads to chocolate that is dull, soft, grainy tasting, and has a melting point a few degrees higher, is thermodynamically more stable [6]. At room temperature the transition from form V to form VI occurs on a timescale of months, and it occurs even faster at elevated temperatures. The chocolate industry expends considerable effort to produce and maintain chocolate in the proper form to ensure a high-quality product with a reasonable shelf life.

Many other areas of chemistry and materials science must cope with molecular crystal polymorphism as well. Crystal packing influences the stability, sensitivity, and detonation characteristics of energetic materials, for instance [7, 8]. It also can have drastic effects on organic semiconductor materials. Solid rubrene currently holds the record for the highest-known carrier mobility in an organic molecular crystal. However, a rubrene derivative with a different crystal packing motif exhibits no measurable carrier mobility [9].

Predicting molecular crystal structure from first principles is extremely challenging. The problem involves (1) a search over many (millions or more) possible crystal packings, (2) the accurate evaluation of the lattice energy of the possible structures (at 0 K), (3) calculation of the finite-temperature thermodynamic contributions, and (4) an understanding of the competition between thermodynamically and kinetically preferred packing arrangements.

Significant progress on the “search” problem has been made in recent years. For example, the Price group uses a hierarchical series of ever-improving theoretical models to screen out unlikely structures and eventually identify the most stable forms [10, 11]. The initial screening might consider some ~107 randomly generated structures with different space group symmetries and numbers of molecules in the asymmetric unit cell using a very simple force field. This simple force field allows one to rule out a significant fraction of the energetically uncompetitive structures. Subsequent rounds improve the quality of the force field, winnowing down the potential structures toward a final prediction. Neumann and co-workers use a mixture of density functional theory (DFT) with dispersion corrections and a force-field fitted to match the DFT results to search over possible structures [1215]. In their work the force field identifies a subset of likely structures which are then refined with DFT. Both strategies proved effective in the two most recent blind tests of crystal structure prediction [16, 17]. Recent progress in crystal structure global optimization algorithms may also help solve the search problem [18].

Once a relatively small number of candidate crystal structures have been identified, discrimination among them requires predicting the lattice energies or relative energies very accurately. This necessitates using a theoretical approach that can handle the subtle balances between intra- and intermolecular interactions that characterize conformational polymorphism [19]. One must treat the diverse non-covalent interactions – hydrogen bonding, electrostatics, induction (a.k.a. polarization), and van der Waals dispersion – with high and uniform accuracy to avoid biasing the predictions toward certain classes of structures (e.g., hydrogen-bonded vs π-stacking motifs). Overall, the energy differences between experimentally observed polymorphs are typically less than 10 kJ/mol, and they are often closer to ~1 kJ/mol.

After obtaining reliable predictions at 0 K, one can start to think about finite temperature effects. The computation of finite-temperature enthalpies, entropies, and free energies is much less mature. The entropic contributions to relative polymorph stabilities are commonly assumed to be smaller than the enthalpic ones [20], but of course there will be many exceptions. Thermal effects are typically estimated using simple (quasi-)harmonic approaches (e.g., [21]) though sampling-based free energy methods are starting to be explored more actively in this context (e.g., [22, 23]).

Finally, real-world crystallization is often driven by kinetics, not thermodynamics. Understanding the kinetics of crystallization is probably even more difficult than the thermodynamics, since it requires a dynamical understanding of the nucleation and crystal growth processes with sufficient accuracy to differentiate correctly among the different packing motifs. Progress in this direction is also being made, but it will likely be quite some time before one can reliably predict which crystal structures will form kinetically under a given set of crystallization conditions.

Nevertheless, the ability to predict the thermodynamically preferred structures reliably would be very useful for predicting temperature and pressure regions of phase stability, to determine whether a given pharmaceutical polymorph is stable or metastable, or to predict a variety of crystal properties. Toward this end, this chapter focuses on computing accurate lattice energies and relative energies at 0 K.

2 Theory

The theoretical treatment of molecular crystals has traditionally relied on either molecular mechanics (MM) force fields or electronic structure methods with periodic boundary conditions. Force field modeling of molecular crystals has improved dramatically over the past decade [19, 2427], as demonstrated by the major successes in the recent blind tests of crystal structure prediction [11, 12, 1517, 28]. Much of this success arises from the inclusion of increasing amounts of quantum mechanical information into the force fields, ranging from the parameterization to the determination of intramolecular conformation.

In that vein, one should be able to achieve even better accuracy by treating the systems fully quantum mechanically. This requires a careful balance between accuracy and computational expense. Most quantum mechanical (QM) calculations on molecular crystals are performed with periodic DFT. Widely used semi-local density functionals generally do not describe van der Waals dispersion interactions. However, there has been substantial progress toward including van der Waals dispersion interactions either self-consistently or as an a posteriori correction using a variety of empirical and non-empirical strategies [2935]. Despite the tremendous progress in this area, it can sometimes be difficult to identify when DFT methods are performing well enough or to interpret the results when different density functionals make contradictory predictions.

Wavefunction methods offer the potential to improve the quality of the predictions systematically by improving the wavefunction. The simplest useful wavefunction technique for molecular crystals is second-order Møller–Plesset perturbation theory (MP2). A number of periodic MP2 implementations exist, including efficient ones based on local-correlation ideas [3647]. These provide a nice alternative to DFT, but they remain relatively computationally expensive. Furthermore, MP2 correlation itself is often insufficient, as will be discussed below. Unfortunately, more accurate periodic coupled cluster implementations are too expensive to be applied to most chemically interesting molecular crystals [4851].

2.1 Fragment-Based Methods in Electronic Structure Theory

Fragment-based methods provide a lower-cost alternative to traditional periodic boundary condition electronic structure methods. These techniques partition the system in some fashion, perform quantum mechanical calculations on each individual fragment, and piece them together to obtain information about the system as a whole. While many individual fragment calculations are needed in a single crystal, each is relatively inexpensive compared to a full periodic crystal calculation. This allows one to utilize higher-level electronic structure methods at much lower computational cost.

Early fragment techniques include the fragment molecular orbital method [52, 53], divide-and-conquer techniques [54], and incremental schemes [55, 56], but there has been an explosion of fragment methods in recent years, as discussed in two reviews [57, 58] and in a thematic issue of Physical Chemistry Chemical Physics [59]. A couple of groups have also identified a common framework that unifies most or all of these fragment methods [60, 61].

One very common fragment strategy decomposes the total energy of a set of interacting molecules, whether in a cluster or a crystal, according to a many-body expansion:

$$ {E}_{\mathrm{total}}={E}_{1\hbox{-} \mathrm{body}}+{E}_{2\hbox{-} \mathrm{body}}+{E}_{3\hbox{-} \mathrm{body}}+\cdots . $$
(1)

The expansion is formally exact, but any computationally useful application of this expansion requires approximation of the higher-order terms in some fashion. For a typical molecular crystal, the three-body and higher terms account for 10–20% of the total interaction energy, making those terms necessary for accurate crystal modeling.

Approximations to the many-body expansion typically fall into two categories. The first category uses electrostatic embedding to incorporate polarization effects into the lower-order terms, thereby reducing the importance of the higher-order terms. The many-body expansion can then be truncated after two-body or three-body terms. Examples of such embedding approaches include the electrostatically-embedded many-body expansion approach [6264], binary interaction [6567], and the exactly embedded density functional many-body expansion [68].

The electrostatic embedding methods are very successful, but they suffer from two potential disadvantages. First, while it is often true that induction effects dominate the many-body contributions, many-body dispersion has also proved important in molecular crystal systems [30, 58, 69, 70], and those effects are not captured via electrostatic embedding. Second, embedding complicates the calculation of nuclear gradients and Hessians of the energy. When embedding a particular monomer or dimer in a potential arising from the other molecules, the embedding potential depends on the positions of all other atoms in the system. Therefore, the energy gradient of that monomer or dimer now depends on all 3N atomic coordinates, instead of just the coordinates of the monomer or dimer in question. These additional gradient contributions arising from the embedding-potential are often neglected, but they can sometimes be significant [65].

The methods in the second category of many-body expansion approximations make no truncation in the many-body expansion. Instead, the higher-order terms are approximated at some lower level of theory. The incremental method and other related schemes treat the important low-order terms with accurate ab initio methods and approximate the higher-order terms using Hartree-Fock (HF), DFT, or any higher-level QM technique. Alternatively, the hybrid many-body interaction (HMBI) approach (and a nearly identical model simultaneously and independently proposed by Manby and co-workers [71]) approximates the higher-order terms using a polarizable force field. The advantage of this approach is that polarizable force fields are much less expensive to evaluate than electronic structure methods. Compared to embedding techniques, these approaches can avoid the need to make a priori assumptions about where to truncate the many-body expansion and provide the flexibility to build in whatever physical terms are needed in the system, including the aforementioned many-body dispersion effects.

2.2 The Hybrid Many-Body Interaction (HMBI) Method

In the HMBI model, the intramolecular (one-body) and short-range (SR) pairwise intermolecular (two-body) interactions are treated with electronic structure theory, while the long-range (LR) two-body and the many-body terms are approximated with the polarizable force field:

$$ {E}_{\mathrm{HMBI}}={E}_{1\hbox{-} \mathrm{body}}^{\mathrm{QM}}+{E}_{\mathrm{SR}2\hbox{-} \mathrm{body}}^{\mathrm{QM}}+{E}_{\mathrm{LR}2\hbox{-} \mathrm{body}}^{\mathrm{MM}}+{E}_{\mathrm{many}\hbox{-} \mathrm{body}}^{\mathrm{MM}}, $$
(2)

where the “many-body” term includes all three-body and higher interactions (Fig. 1). To evaluate the energy in practice, one exploits the fact that one can write a many-body expansion purely in terms of MM contributions:

Fig. 1
figure 1

Schematic of the HMBI method. Each individual molecule in the unit cell and its short-range pairwise interactions are treated with QM, while longer-range interactions and interactions involving more than two molecules are treated with MM. The shaded region indicates the smooth transition QM and MM via interpolation. Reprinted with permission from [115]. Copyright 2010 American Chemical Society

$$ {E}_{\mathrm{total}}^{\mathrm{MM}}={E}_{1\hbox{-} \mathrm{body}}^{\mathrm{MM}}+{E}_{\mathrm{SR}2\hbox{-} \mathrm{body}}^{\mathrm{MM}}+{E}_{\mathrm{LR}2\hbox{-} \mathrm{body}}^{\mathrm{MM}}+{E}_{\mathrm{many}\hbox{-} \mathrm{body}}^{\mathrm{MM}}, $$
(3)

which can be rearranged as

$$ {E}_{\mathrm{LR}2\hbox{-} \mathrm{body}}^{\mathrm{MM}}+{E}_{\mathrm{many}\hbox{-} \mathrm{body}}^{\mathrm{MM}}={E}_{\mathrm{total}}^{\mathrm{MM}}-{E}_{1\hbox{-} \mathrm{body}}^{\mathrm{MM}}-{E}_{\mathrm{SR}2\hbox{-} \mathrm{body}}^{\mathrm{MM}}. $$
(4)

Substituting this expression into (2) leads to the following working HMBI energy expression:

$$ {E}_{\mathrm{total}}^{\mathrm{HMBI}}={E}_{\mathrm{total}}^{\mathrm{MM}}+{\displaystyle \sum_i}\;\left({E}_i^{\mathrm{QM}}-{E}_i^{\mathrm{MM}}\right)+{\displaystyle \sum_{ij}}\;{d}_{ij}(R)\left({\Delta}^2{E}_{ij}^{\mathrm{QM}}-{\Delta}^2{E}_{ij}^{\mathrm{MM}}\right), $$
(5)

where E i corresponds to the energy of monomer i, and Δ2 E ij  = E ij  − E i  − E j is the interaction energy between monomers i and j evaluated as the difference between the total energy of the dimer E ij and the individual monomer energies E i and E j . Note that the monomer energies in these expressions are evaluated at the same geometry as in the dimer (and in the full cluster or crystal). The d ij (R) term ensures no discontinuities arise in the potential energy surface as the model transitions from the short-range QM to the long-range MM two-body interaction regimes. This function decays from 1 to 0 as a function of the intermolecular distance R (defined here as the shortest distance between the two molecules) over a user-defined interval governed by two parameters, r 1 and r 0 [72]:

$$ {d}_{ij}(R)=\frac{1}{1+{e}^{2\left|{r}_1-{r}_0\right|/\left({r}_1-R\right)-\left|{r}_1-{r}_0\right|/\left(R-{r}_0\right)}}. $$
(6)

By default, we conservatively transition from QM to MM between r 1 = 9.0 Å and r 0 = 10.0 Å, but more aggressive cutoffs can often be used. In water/ice, for instance, using the ab initio force field described in the next section, one can transition from QM to MM between 4.5 and 5.5 Å with virtually no loss in accuracy (see Sect. 2.3).

The energy expression in (5) resembles the expressions used in a two-layer ONIOM-style QM/MM model. One computes the energy of the entire system at a low-level of theory, and then corrects it with smaller calculations at a higher level of theory. The physics described by the two types of models is quite different, however. In ONIOM QM/MM, one partitions the system into distinct QM and MM regions. In the fragment-based HMBI QM/MM approach, one instead partitions based on the nature of the interaction. There are no specific QM and MM regions. Each molecule has both QM and MM interactions in the HMBI model. The important interactions are treated with QM, while the less important ones are treated with MM. In this sense, the HMBI model is spatially homogeneous, as is appropriate for modeling a molecular crystal.

For systems with periodic boundary conditions like crystals, the HMBI energy expression in (5) must also include pairwise interactions between central unit cell molecules and their periodic images:

$$ \begin{array}{c}{E}_{\mathrm{total}}^{\mathrm{HMBI}}={E}_{\mathrm{total}}^{\mathrm{MM}}+{\displaystyle \sum_i}\;\left({E}_i^{\mathrm{QM}}-{E}_i^{\mathrm{MM}}\right)+{\displaystyle \sum_{ij}}\;{d}_{ij(0)}\left({\Delta}^2{E}_{ij(0)}^{\mathrm{QM}}-{\Delta}^2{E}_{ij(0)}^{\mathrm{MM}}\right)\\ {}+\frac{1}{2}{\displaystyle \sum_i}\;{\displaystyle \sum_{j\left(\mathbf{n}\right)}^{\mathrm{images}}}\;{d}_{ij\left(\mathbf{n}\right)}\left({\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}-{\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}\right).\end{array} $$
(7)

In this expression, i runs over molecules in the central unit cell, while j can either be in the central unit cell (j(0)) or in the periodic image cell n (j(n)). Unit cell n is defined as the cell whose origin lies at vector \( \mathbf{n}={n}_{v_1}{\mathbf{v}}_1+{n}_{v_2}{\mathbf{v}}_2+{n}_{v_3}{\mathbf{v}}_3 \) in terms of the three lattice vectors, v 1, v 2, and v 3. In practice, thanks to the function d ij (R), the sum over j(n) runs only over molecules within a distance r 0 of the current central unit cell molecule i.

The computational bottleneck in HMBI is the evaluation of the QM pairwise interaction energies Δ2 E ij(0) and Δ2 E ij(n). Because only short-range pairwise interactions are treated with QM, the HMBI model scales linearly with the number of molecules in the system (non-periodic) or unit-cell (periodic). More precisely, the MM terms do not scale linearly, but their cost is so much smaller than that of the QM terms for any practical system that linear scaling behavior is observed. For non-periodic systems, the onset of linear scaling occurs once the system becomes larger than the outer extent of the QM-to-MM transition region, r 0. In periodic systems, which are formally infinite, linear-scaling behavior is observed for unit cells of any size. This linear-scaling behavior is a key advantage of fragment methods compared to fully QM methods like periodic DFT or MP2. Large unit cells or supercells of the sort that might be used to perform lattice dynamics or to examine localized/defect behavior can be much cheaper with a fragment method than with a traditional periodic QM methods.

2.2.1 Nuclear Gradients and Hessians

The HMBI energy expression contains only additive energy contributions and does not use any sort of embedding, so derivatives of the energy can be computed straightforwardly. For instance, if q l corresponds to the Cartesian x, y, or z coordinates (not its fractional coordinate) of the l-th atom in the central unit cell, then the gradient of the energy with respect to q l is given by [73]

$$ \begin{array}{c}\frac{\partial {E}_{\mathrm{total}}^{\mathrm{HMBI}}}{\partial {q}_l}=\frac{\partial {E}_{\mathrm{total}}^{\mathrm{MM}}}{\partial {q}_l}+{\displaystyle \sum_i}\;\left(\frac{\partial {E}_i^{\mathrm{QM}}}{\partial {q}_l}-\frac{\partial {E}_i^{\mathrm{MM}}}{\partial {q}_l}\right)+{\displaystyle \sum_{ij\left(\mathbf{0}\right)}}\;{\scriptscriptstyle \frac{\partial {d}_{ij(0)}}{\partial {q}_l}}\left({\Delta}^2{E}_{ij(0)}^{\mathrm{QM}}-{\Delta}^2{E}_{ij(0)}^{\mathrm{MM}}\right)\\ {}+{\displaystyle \sum_{ij\left(\mathbf{0}\right)}}\;{d}_{ij(0)}\left(\frac{\partial {\Delta}^2{E}_{ij(0)}^{\mathrm{QM}}}{\partial {q}_l}-\frac{\partial {\Delta}^2{E}_{ij(0)}^{\mathrm{MM}}}{\partial {q}_l}\right)\\ {}+{\scriptscriptstyle \frac{1}{2}}{\displaystyle \sum_i}\;{\displaystyle \sum_{j\left(\mathbf{n}\right)}^{\mathrm{images}}}\;\frac{\partial {d}_{ij\left(\mathbf{n}\right)}}{\partial {q}_l}\left({\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}-{\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}\right)\\ {}+{\scriptscriptstyle \frac{1}{2}}{\displaystyle \sum_i}\;{\displaystyle \sum_{j\left(\mathbf{n}\right)}^{\mathrm{images}}}\;{d}_{ij\left(\mathbf{n}\right)}\left(\frac{\partial {\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}}{\partial {q}_l}-\frac{\partial {\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}}{\partial {q}_l}\right).\end{array} $$
(8)

The individual one-body and two-body energy gradient terms in (8) are obtained readily from the monomer and dimer gradients computed in standard electronic structure or MM software packages. The expression for the gradient of the QM-to-MM smoothing function d ij has been provided previously [73].

To optimize the size and shape of the unit cell, one also needs the gradient with respect to the lattice vectors. When working in Cartesian coordinates instead of fractional coordinates, changing the lattice vectors does not affect the one-body or two-body terms within the central unit cell, but it does affect the two-body contributions due to interactions between molecules in the central unit cell and those in periodic image cells and the MM energy of the entire crystal. The resulting gradient with respect to the qth coordinate (x, y, or z) of the ϵth lattice vector (v 1, v 2, or v 3) is given by

$$ \begin{array}{c}\frac{\partial E}{\partial {v}_{\upepsilon q}}=\frac{\partial {E}_{\mathrm{total}}^{\mathrm{MM}}}{\partial {v}_{\upepsilon q}}+\frac{1}{2}{\displaystyle \sum_i}\;{\displaystyle \sum_{j\left({n}_v\right)}}\;{n}_v{\displaystyle \sum_k}\;\left\{\frac{\partial {d}_{ij\left({n}_v\right)}}{\partial {q}_k}\left({\Delta}^2{E}_{ij\left({n}_v\right)}^{\mathrm{QM}}-{\Delta}^2{E}_{ij\left({n}_v\right)}^{\mathrm{MM}}\right)\right.\\ {}\left.+{d}_{ij\left({n}_v\right)}\left(\left(\frac{\partial {E}_{ij\left({n}_v\right)}^{\mathrm{QM}}}{\partial {q}_k}-\frac{\partial {E}_{j\left({n}_v\right)}^{\mathrm{QM}}}{\partial {q}_k}\right)-\left(\frac{\partial {E}_{ij\left({n}_v\right)}^{\mathrm{MM}}}{\partial {q}_k}-\frac{\partial {E}_{j\left({n}_v\right)}^{\mathrm{MM}}}{\partial {q}_k}\right)\right)\right\}.\end{array} $$
(9)

where k sums over the atoms in periodic image monomer j. Often, one expresses the unit cell in terms of three lattice constants (a, b, and c) and three angles (α, β, and γ). The expressions for the nine components of the lattice vector gradients in (9) can be transformed into expressions for these six lattice parameters [73].

Note that, except for the term ∂E MMtotal /∂v ϵq , all the terms needed to compute the lattice vector gradients in (9) are already available from the gradients with respect to the atomic positions (8). The ∂E MMtotal /∂v ϵq term can be computed inexpensively, since it is at the MM level. Therefore, the calculation of the lattice vector gradient typically requires minimal additional work once the nuclear gradients with respect to atomic position have been obtained. The evaluation of the one-body and two-body QM gradients forms the computational bottleneck in evaluating the full nuclear gradient.

The nuclear Hessian can be computed similarly by differentiating (8) with respect to a second nuclear coordinate, \( {q}_{l^{\prime }} \):

$$ \begin{array}{c}\frac{\partial^2{E}_{\mathrm{total}}^{\mathrm{HMBI}}}{\partial {q}_l\partial {q}_{l^{\prime }}}=\frac{\partial {E}_{\mathrm{PBC}}^{\mathrm{MM}}}{\partial {q}_l\partial {q}_{l^{\prime }}}+{\displaystyle \sum_i}\;\left(\frac{\partial^2{E}_i^{\mathrm{QM}}}{\partial {q}_l\partial {q}_{l^{\prime }}}-\frac{\partial^2{E}_i^{\mathrm{MM}}}{\partial {q}_l\partial {q}_{l^{\prime }}}\right)\\ {}+{\displaystyle \sum_{ij\left(\mathbf{n}\right)}}\;{\zeta}_{ij}{d}_{ij\left(\mathbf{n}\right)}\left(\frac{\partial^2{\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}}{\partial {q}_l\partial {q}_{l^{\prime }}}-\frac{\partial^2{\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}}{\partial {q}_l\partial {q}_{l^{\prime }}}\right)\\ {}+{\displaystyle \sum_{ij\left(\mathbf{n}\right)}}\;{\zeta}_{ij}\frac{\partial {d}_{ij\left(\mathbf{n}\right)}}{\partial {q}_l}\left(\frac{\partial {\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}}{\partial {q}_{l^{\prime }}}-\frac{\partial {\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}}{\partial {q}_{l^{\prime }}}\right)\\ {}+{\displaystyle \sum_{ij\left(\mathbf{n}\right)}}\;{\zeta}_{ij}\frac{\partial {d}_{ij\left(\mathbf{n}\right)}}{\partial {q}_{l^{\prime }}}\left(\frac{\partial {\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}}{\partial {q}_l}-\frac{\partial {\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}}{\partial {q}_l}\right)\\ {}+{\displaystyle \sum_{ij\left(\mathbf{n}\right)}}\;{\zeta}_{ij}\frac{\partial^2{d}_{ij\left(\mathbf{n}\right)}}{\partial {q}_l\partial {q}_{l^{\prime }}}\left({\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{QM}}-{\Delta}^2{E}_{ij\left(\mathbf{n}\right)}^{\mathrm{MM}}\right),\end{array} $$
(10)

where the sums over j(n) run over molecules in both the central unit cell and in periodic image cells, ζ ij  = 1 for dimers where both monomers lie within the central unit cell (n = 0), and ζ ij  = 1/2 for dimers where the second monomer lies outside the central unit cell. Evaluation of the Hessian requires gradients and Hessians for each individual monomer and dimer. Once the Hessian has been computed, one can compute harmonic vibrational frequencies, lattice dynamics, statistical thermodynamic partition functions, etc.

2.2.2 Crystal Symmetry

Molecular crystals often exhibit high symmetry, both translational (due to the periodic boundary conditions) and space group symmetry, and significant computational savings can be reaped by exploiting this symmetry in the monomer and dimer calculations. The straightforward approach would identify the space group of the crystal and use the symmetry operations of that group to determine which monomers and dimers are symmetry-equivalent. One need only compute the energies and forces for the symmetry-unique monomers and dimers and then scale their contributions based on the number of symmetry-equivalent monomers or dimers.

Alternatively, one can simply rotate the monomers and dimers to a common reference frame (e.g., aligned along the principle axes of inertia) and simply test which monomer/dimer geometries are identical within some numerical threshold. This approach avoids the need (1) to specify the space group and (2) to program the symmetry operations for all 230 space groups.

A sizable fraction of organic molecular crystals exhibit P21/c symmetry, for which roughly fourfold computational savings can be obtained. The savings can be even larger for other space groups. For acetamide crystals in the R3c space group, one obtains 18-fold speed-ups by exploiting symmetry. So while the details are system-dependent, for a crystal with one molecule in the asymmetric unit cell and a conservative QM to MM transition distance, one might typically need to perform ~50 symmetry-unique dimer calculations.

2.3 Accurate Force Fields for Long-Range and Many-Body Interactions

The success of the HMBI approach depends critically on the quality of the polarizable force field used for the long-range two-body and the many-body terms; c.f. (2). This means that it needs to capture two-body electrostatics, two-body van der Waals dispersion, self-consistent long-range and many-body induction, and many-body dispersion interactions (approximated here with only the leading three-body Axilrod–Teller term):

$$ {E}^{\mathrm{MM}}={E}_{2\hbox{-} \mathrm{bodyes}}+{E}_{2\hbox{-} \mathrm{bodydisp}}+{E}_{\mathrm{induction}}+{E}_{3\hbox{-} \mathrm{bodydisp}}. $$
(11)

In our early work [74] we used the Amoeba force field [75, 76], which includes all of these contributions except for the many-body dispersion. It works fairly well in this context, but even better results are achieved by constructing an ab initio force field (AIFF) “on the fly” based on QM calculations for each individual monomer [77, 78].

The key idea behind the AIFF is to parameterize the force field in terms of atom-centered distributed multipole moments [7981], distributed polarizabilities [82, 83], and distributed dispersion coefficients [84]. These are obtained from the molecular electron density, the static polarizabilities, and the frequency-dependent polarizabilities, respectively. The form of the force field is well-justified at long ranges from intermolecular perturbation theory, and empirical short-range damping functions help avoid serious problems for shorter-range interactions [85, 86].

This AIFF model mimics the much more expensive QM treatment very well, as shown in Figs. 2, 3, and 4. The use of multipolar expansions and the lack of exchange terms lead to some problems at short range, but the long-range interactions are modeled very accurately. Figure 2 shows that the predicted lattice energy of ice Ih is virtually invariant to the distance at which HMBI transitions from a QM to an MM description of the long-range pairwise interactions once the two molecules are ~4.5 Å apart. For other systems, the corresponding transition distance may need to be longer (e.g., ~7 Å in formamide), but the AIFF always behaves well at sufficiently long distances [58].

Fig. 2
figure 2

Performance of the AIFF for long-range pairwise interactions: Change in the lattice energy of ice Ih as a function of the QM to MM transition distance (r 0 in (6), and r 1 = r 0 − 1 Å). Starting around ~4.5 Å, the AIFF reproduces the QM to within 0.1 kJ/mol. Neglecting the MM terms entirely leads to much larger errors [58]. Adapted with permission of the PCCP Owner Societies

Fig. 3
figure 3

Performance of the AIFF for many-body induction: error distributions for the AIFF and Amoeba force field many-body induction relative to MP2 for 101 (formamide)8 geometries. The AIFF errors are smaller than the Amoeba ones and are nearly centered on zero. Adapted with permission from [77]. Copyright 2010 American Chemical Society

Fig. 4
figure 4

Performance of the AIFF for three-body dispersion: Correlation between the HMBI three-body dispersion and the same contribution evaluated from symmetry adapted perturbation theory (SAPT) for 78 trimers taken from the formamide crystal [58]. Adapted with permission of the PCCP Owner Societies

The AIFF also reproduces the QM many-body induction effects accurately [87]. Figure 3 shows the errors in the many-body induction in a set of 101 (formamide)8 geometries for the Amoeba force field and AIFF relative to RI-MP2. The Amoeba force field performs fairly well, but it systematically underestimates the many-body induction interactions by up to ~2 kJ/mol. The AIFF performs significantly better, with errors of roughly ±0.5 kJ/mol and a mean error very close to zero. Compared to Amoeba, the success of the AIFF stems from (1) its use of higher-rank multipoles (up to hexadecapole), (2) its use of higher-rank polarizabilities (up to quadrupole–quadrupole), and (3) the fact that these parameters are computed for each molecule in its current geometry rather than frozen at the values for some averaged/equilibrium geometry.

Finally, Fig. 4 demonstrates that the three-body dispersion model performs well compared to the three-body dispersion term in symmetry-adapted perturbation theory (SAPT). The simple, isotropic coupled Kohn–Sham dispersion coefficient representation provides a good approximation to the more complete SAPT calculation [58].

Overall, the AIFF does an excellent job of reproducing the QM interactions it replaces at much lower cost. HMBI predictions for the lattice energies of ammonia and carbon dioxide crystals differ from full periodic MP2 by only 1–2 kJ/mol, for instance [58].

2.3.1 AIFF Implementation

The long-range two-body electrostatics in the AIFF are implemented in standard fashion, with electrostatic interaction energy between two atoms A and B being given by

$$ {E}_{\mathrm{es}}\leftarrow {\displaystyle \sum_{tu}}\;{Q}_t^{\mathrm{A}}{T}_{tu}{Q}_u^{\mathrm{B}}, $$
(12)

where the T tu matrix includes the distance- and orientation-dependent contributions for the interaction of two different spherical-tensor multipole moment components Q t and Q u . To evaluate the induction contributions, one first finds the induced multipole moments according to

$$ \Delta {Q}_t^{\mathrm{A}}=-{\displaystyle \sum_{t^{\prime }u}}\;{\alpha}_{t{t}^{\prime}}^{\mathrm{A}}{T}_{t^{\prime }u}\left({Q}_u^{\mathrm{B}}+\Delta {Q}_u^{\mathrm{B}}\right), $$
(13)

where \( {\alpha}_{t{t}^{\prime}}^{\mathrm{A}} \) is the static polarizability tensor on atom A and ΔQ is an induced multipole moment. Clearly the induced multipole moment on atom A depends on the induced multipole moment on atom B, so this process is done self-consistently until the induced multipoles reach convergence. Once the induced multipoles are known, one can compute the induction energy contribution for the pair of atoms A and B according to

$$ {E}_{\mathrm{ind}}\leftarrow {\displaystyle \sum_{tu}}\;\Delta {Q}_t^{\mathrm{A}}{T}_{tu}{Q}_u^{\mathrm{B}}+{Q}_t^{\mathrm{A}}{T}_{tu}\Delta {Q}_u^{\mathrm{B}}. $$
(14)

The two-body dispersion between atoms A and B is evaluated via Casimir–Polder integration over the frequency-dependent polarizabilities:

$$ {E}_{2\hbox{-} \mathrm{bodydisp}}\leftarrow {\displaystyle \sum_{tu}}\;{\displaystyle \sum_{t^{\prime }{u}^{\prime }}}\;{T}_{tu}{T}_{t^{\prime }{u}^{\prime }}{\displaystyle {\int}_0^{\infty }}\;{\alpha}_{t{t}^{\prime}}^{\mathrm{A}}\left( i\omega \right){\alpha}_{u{u}^{\prime}}^{\mathrm{B}}\left( i\omega \right) d\omega . $$
(15)

The integration is typically evaluated via numerical quadrature at ten frequencies ω. The expression in (15) corresponds to an anisotropic model for atom-atom dispersion. However, to a fairly good approximation, one can approximate this with a simple isotropic dispersion model (i.e., one which averages over the diagonal dipole–dipole and quadrupole–quadrupole elements of the frequency-dependent polarizability and neglects the off-diagonal elements). In that case, the dispersion model reduces to the standard C 6, C 8, etc., terms divided by the interatomic distance R to the corresponding power,

$$ {E}_{2\hbox{-} \mathrm{bodydisp}}\leftarrow \frac{C_6^{\mathrm{AB}}}{R_{\mathrm{AB}}^6}+\frac{C_8^{\mathrm{AB}}}{R_{\mathrm{AB}}^8}+\cdots . $$
(16)

The isotropic dispersion coefficients C n are obtained from the Casimir–Polder integration over the appropriate elements of the isotropic frequency-dependent polarizabilities \( \overline{\alpha} \):

$$ {C}_n^{\mathrm{A}\mathrm{B}}\propto {\displaystyle {\int}_0^{\infty }}\;{\displaystyle {\overline{\alpha}}^{\mathrm{A}}}\left( i\omega \right){\displaystyle {\overline{\alpha}}^{\mathrm{B}}}\left( i\omega \right) d\omega . $$
(17)

The many-body dispersion is approximated using the leading Axilrod–Teller three-body dispersion contribution. In the isotropic case, the three-body dispersion between atoms A, B, and C is computed according to

$$ {E}_{3\hbox{-} \mathrm{bodydisp}}\leftarrow {C}_9^{\mathrm{AB}\mathrm{C}}\frac{\left(1+3 \cos \widehat{A} \cos \widehat{B} \cos \widehat{C}\right)}{R_{\mathrm{AB}}^3{R}_{\mathrm{BC}}^3{R}_{\mathrm{AC}}^3}, $$
(18)

where the R IJ correspond to the distances between pairs of atoms and the cosines involve the interior angles of the triangle formed by the three atoms. The C 9 dispersion coefficient is obtained via Casimir–Polder integration over the frequency-dependent dipole–dipole polarizabilities on all three atoms:

$$ {C}_9^{\mathrm{A}\mathrm{BC}}\propto {\displaystyle {\int}_0^{\infty }}\;{\displaystyle {\overline{\alpha}}^{\mathrm{A}}}\left( i\omega \right){\displaystyle {\overline{\alpha}}^{\mathrm{B}}}\left( i\omega \right){\displaystyle {\overline{\alpha}}^{\mathrm{C}}}\left( i\omega \right) d\omega . $$
(19)

Note that most of the interaction terms described here also involve various empirical damping functions at short range to avoid divergences and the “polarization catastrophe,” [78] but those are omitted here for simplicity.

The most difficult aspect of the force field implementation involves the evaluation of the E MMtotal term in (7). The periodic crystal’s induced multipoles are first computed self-consistently in a large, finite cluster consisting of the central unit cell and some number of periodic images (e.g., those within 25 Å). The algorithm works by inducing multipoles on the central unit cell molecules, replicating those induced moments on the periodic image molecules, and repeating the process until self-consistency is achieved. This procedure minimizes the edge effects arising from the finite cluster. Next, the permanent and induced multipole moments are combined and the overall interaction is evaluated via a multipolar Ewald sum [88]. Finally, the two-body and three-body dispersion contributions are evaluated via explicit lattice summation with large cutoffs. The evaluation of the AIFF interactions is cheap compared to the QM calculations, so one can afford to evaluate all of these contributions with tight cutoffs, thereby minimizing any errors due to truncating the lattice sums. See [78] for more details on the AIFF implementation and its performance for molecular crystals.

2.4 Electronic Structure Treatment of the Intermolecular Interactions

As the results described in Sect. 1 will demonstrate, high quality electronic structure methods need to be used for the QM one-body and two-body terms in HMBI. The electronic structure treatment must be capable of balancing the different types of intramolecular and intermolecular interactions to discriminate properly among different packing motifs. Ideally, one would employ high-level coupled cluster methods, like coupled cluster singles, doubles, and perturbative triples (CCSD(T)), with large basis sets, but this is usually impractical due to its N 7 scaling with system size N.

A more pragmatic approach will primarily use techniques that scale no more than N 5, like MP2. It is well known, however, that while MP2 describes van der Waals dispersion, it does so poorly [8991]. Comparisons with intermolecular perturbation theory reveal the reasons for this poor performance: MP2 treats intermolecular dispersion interactions at the uncoupled HF (UCHF) level [92, 93]. In this view, the intermolecular dispersion interaction involves matrix elements of the intermolecular interaction between the ground state and excited state wavefunctions for molecules A and B divided by an energy denominator that depends on the excitation energies from ground to excited states:

$$ {E}_{\mathrm{disp}}=-{\displaystyle \sum_{ab}}\;\frac{<{\psi}_{\mathrm{A}}^0{\psi}_{\mathrm{B}}^0\left|\widehat{V}\right|{\psi}_{\mathrm{A}}^a{\psi}_{\mathrm{B}}^b>}{E_{\mathrm{A}}^a-{E}_{\mathrm{A}}^0+{E}_{\mathrm{B}}^b-{E}_{\mathrm{B}}^0}. $$
(20)

In UCHF (and MP2) these excited states for ψ a and excitation energies E a − E 0 are approximated using unrelaxed, ground-state HF orbitals and Fock matrix orbital eigenvalue energy differences.

Better results can be obtained by replacing the UCHF treatment of intermolecular dispersion with a coupled HF (CHF) [94] or coupled Kohn–Sham (CKS) [95, 96] one in which the excited states and excitation energies are computed with time-dependent HF or time-dependent DFT. Or one can obtain improved dispersion coefficients through other means [97]. The CKS variant has proved particularly successful, and is known as MP2C:

$$ {E}_{\mathrm{MP}2\mathrm{C}}={E}_{\mathrm{MP}2}-{E}_{\mathrm{disp}}^{\mathrm{UCHF}}+{E}_{\mathrm{disp}}^{\mathrm{CKS}}. $$
(21)

MP2C works very well for a variety of intermolecular interaction types across a wide spectrum of intermolecular separations and orientations [89, 91, 96, 98102]. The dispersion correction scales N 4, so MP2C retains the overall fifth-order scaling of MP2. However, the prefactor for the dispersion correction is relatively large, and it consumes a non-trivial amount of computer time. For example, computing the HMBI RI-MP2C/aug-cc-pVTZ single-point energy of the aspirin crystal requires ~2,450 h for the MP2 and ~390 h for the dispersion correction. So while the underlying MP2 calculation clearly dominates, accelerating the MP2C dispersion correction would be beneficial.

MP2C appears to work very well for molecular crystal problems [70, 103]. Of course, other alternatives for accurately describing intermolecular interactions with similar computational cost exist, including spin-component-scaled MP2 methods [104106], dispersion-weighted MP2 [107], and van der Waals-corrected density functionals. Recent improvements in the random-phase approximation (RPA) are also promising and may prove useful in the near future [108111].

2.4.1 Faster MP2C for Molecular Crystals

Evaluating the MP2C dispersion correction requires computing UCHF and CKS density–density-response functions χ for each monomer at a series of frequencies ω,

$$ {\displaystyle {\left[{\chi}_0\left(\omega \right)\right]}_{PQ}}=-4{\displaystyle \sum_{ia}}\;\frac{\left(P\Big| ia\right){\upepsilon}_{ia}\left( ia\Big|Q\right)}{\upepsilon_{ia}^2+{\omega}^2}, $$
(22)

and then computing the dispersion energy via Casimir–Polder integration [96]:

$$ {E}_{\mathrm{disp}}=-\frac{1}{2\pi }{\displaystyle {\int}_0^{\infty }}\; d\omega {\displaystyle {\tilde{\chi}}^{\mathrm{A}}}\left(\omega \right){\mathbf{J}}^{\mathrm{A}\mathrm{B}}{\left({\displaystyle {\tilde{\chi}}^{\mathrm{B}}}\left(\omega \right)\right)}^T{\left({\mathbf{J}}^{\mathrm{A}\mathrm{B}}\right)}^T. $$
(23)

In these expressions, i and a are occupied and virtual molecular orbitals, P and Q are auxiliary basis functions, ϵ ia is the difference between the HF orbital energies for orbitals i and a, J AB = (P A|r − 112 |Q B), \( \tilde{\chi}={\mathbf{S}}^{-1}\chi {\mathbf{S}}^{-1} \), and S = (P|r − 112 |Q). For CKS, one must obtain the coupled density response functions from the uncoupled one in (22) by solving the Dyson equation [96].

Equations (22) and (23) are typically computed using a so-called dimer-centered (DC) basis, in which the calculations on the isolated monomers are carried out in the presence of ghost basis functions on the other monomer it is interacting with. This has two disadvantages for fragment-based calculations in molecular crystals. First, the ghost basis functions substantially increase the basis set size and therefore the computational cost. Second, when computing the many different pairwise intermolecular interaction energies, one must recompute each monomer’s response function repeatedly for each different dimer interaction in which it is involved.

However, while the calculation of accurate CKS or UCHF dispersion energies requires large basis sets [112], it turns out that the energy difference between them is much less sensitive to the basis set [103]. In fact, one obtains nearly identical results when using a comparable monomer-centered (MC) basis (i.e., with no ghost basis functions). This indicates that the basis-set dependence of UCHF and CKS dispersion are similar, and the higher-order basis-set effects are well described in the DC UCHF dispersion that is inherent in the supermolecular MP2 calculation. It is unnecessary to replace those contributions with CKS ones. Figure 5 demonstrates the excellent agreement between the MP2C results obtained in monomer-centered and dimer-centered basis sets, nearly independent of whether one uses a double-zeta basis or extrapolated complete-basis-set limit.

Fig. 5
figure 5

For the dimers in the S22 test set [139], the MP2C dispersion correction calculated in an MC basis is nearly identical to that computed in a DC basis, even for a small aug-cc-pVDZ basis. Adapted with permission from [103]. Copyright 2013, American Institute of Physics

In practice, switching to an MC basis accelerates the calculation of the dispersion correction for a single dimer roughly fivefold. Much more dramatic savings can be obtained in a molecular crystal, however. Thanks to space group symmetry and the use of periodic boundary conditions, typical molecular crystal unit cells contain no more than a handful of symmetry-unique monomers. Therefore, one needs only calculate the monomer density–density response functions in (22) for those few unique monomers. Then one can compute the dispersion interaction for each dimer according to (23) with trivial effort. For instance, using this approach reduces the computational time for evaluating the MP2C/aug-cc-pVTZ dispersion correction in crystalline aspirin by two orders of magnitude, from 390 CPU hours to only 2.8 CPU hours. At the same time, the MC MP2C approach affects the relative energies of the two polymorphs by no more than a few tenths of a kJ/mol relative to the DC MP2C values (e.g., Fig. 6). For all practical purposes, this approach makes computing the MP2C correction “free” for molecular crystals. With its high accuracy and low cost, MP2C makes an excellent choice for describing the non-covalent interactions in molecular crystals.

Fig. 6
figure 6

Comparison between MC MP2C and DC MP2C for the relative energies of the five experimentally known polymorphs of oxalyl dihydrazide. The relative polymorph energies differ by 0.3 kJ/mol or less. Adapted with permission from [103]. Copyright 2013, American Institute of Physics

Of course, the aforementioned speed-ups only affect the evaluation of the dispersion correction. As noted previously, the underlying RI-MP2/aug-cc-pVTZ calculations in aspirin require ~2,450 CPU hours. Further computational savings must come from the MP2 part such as through local MP2 or other similar methods. One must be careful, however, that the truncation schemes do not hinder the ability to describe the non-covalent interactions at intermediate distances (i.e., beyond nearest-neighbor interactions but not yet far enough apart to be treated by the force field in HMBI).

2.4.2 Basis Sets

It is well know that describing intermolecular interactions with correlated wave function methods requires large basis sets, so it is no surprise that molecular crystal energetics can often be very sensitive to the basis sets. This is particularly true for cases of conformational polymorphism, where the crystal packing motifs differ in both intramolecular conformation and intermolecular packing. Counterpoise corrections can help correct for basis set superposition error (BSSE). Different conformational polymorphs, however, may exhibit varying degrees of intramolecular BSSE, which is much harder to correct. In the oxalyl dihydrazide example discussed in Sect. 3.3, for example, the qualitative ordering of the experimental polymorphs differs dramatically as a function of basis set for exactly this reason. The intramolecular conformations of certain functional groups, such as the pyramidalization of nitrogen groups, can also be sensitive to basis sets.

In general, one should take care to ensure that molecular crystal calculations are converged with respect to basis set. Basis set extrapolation toward the complete basis set (CBS) limit, where feasible, can be very useful. One can also employ explicitly correlated techniques like MP2-F12 to achieve large-basis accuracy at reasonable computational cost [113, 114].

3 Performance and Applications of HMBI

To demonstrate the capabilities of the HMBI method, this section discusses molecular crystal benchmarks for predicting crystal geometries and lattice energies (Sect. 3.1) along with applications to polymorphic aspirin (Sect. 3.2) and oxalyl dihydrazide (Sect. 3.3) crystals.

The ability to predict crystal lattice energies accurately provides a demanding test for any theoretical treatment of molecular crystals. Lattice energies measure the energy required to dissociate the crystal to isolated gas-phase molecules. Whereas relative polymorph energies allow for some degree of error cancellation between the treatments of two different sets of non-covalent interactions, lattice energies expose any errors in calculating the strength of those interactions (at least for small, rigid molecules for which the intramolecular geometry is similar in the gas and crystal phases). The situation is analogous to the differences between computing reaction energies vs atomization energies, with the former being much easier due to cancellation of errors. In molecular crystals, the degree of error cancellation in the relative energies between polymorphs will depend on the differences in intramolecular configurations and intermolecular packing. In cases like aspirin, where the two polymorphs exhibit similar structures, substantial error cancellation occurs. On the other hand, cases like oxalyl dihydrazide exhibit more diverse interactions and packing modes, leading to less error cancellation and making it much more difficult to obtain accurate polymorphic energies.

3.1 Predicting Molecular Crystal Lattice Energies and Geometries

3.1.1 Lattice Energies

Benchmark lattice energy predictions have been performed on a series of seven small-molecule crystals: ammonia, acetamide, benzene, carbon dioxide, formamide, ice Ih, and imidazole. These crystals were chosen because they span a diverse range of intermolecular interactions, ranging from hydrogen-bonded to van der Waals dispersion bonded. The calculations start in a relatively small basis (aug-cc-pVDZ) RI-MP2, then increase the basis set toward the TQ-extrapolated basis set limit, and then finally examine higher-order correlation effects evaluated with CCSD(T) in a modest basis. Initially the Amoeba force field was used for the MM terms [115], but even better results are obtained when the AIFF (in the Sadlej basis set [116, 117]) described in Sect. 2.3 is used [78].

Fig. 7
figure 7

Benchmark lattice energy predictions compared to experiment (in kJ/mol). Improving the QM treatment systematically improves the lattice energy predictions. The shaded region highlights ±2 kJ/mol to indicate the excellent agreement with the nominal experimental values. See [78] for complete details. Adapted with permission from [78]. Copyright 2011 American Chemical Society

The primarily hydrogen-bonded crystals exhibit generally consistent improvement toward the experimental lattice energies as the electronic structure treatment of the one-body and short-range two-body terms is improved, as shown in Fig. 7. For those cases, the higher-order CCSD(T) correlation contributions are generally small. However, for the benzene and imidazole crystals, where the π-electron van der Waals dispersion interactions are significant, MP2 over-binds the crystals by ~10–12 kJ/mol (~10–20%). This sort of error is particularly problematic in the context of molecular crystal polymorphism, since MP2 will be biased toward such dispersion-bound packing motifs over hydrogen-bonded ones.

The CCSD(T) results correct the MP2 overestimate of the dispersion interactions in these two crystals. Overall, for six of the seven crystals, the estimated CBS-limit CCSD(T) results lie within 1–2 kJ/mol of the nominal experimental values, which is likely within the experimental errors. The reason for the larger errors in acetamide is unclear. Of course, CCSD(T) calculations will often be cost-prohibitive in molecular crystals. MP2C can provide a lower-cost alternative that substantially improves upon MP2. Table 1 demonstrates that including the MP2C dispersion correction has relatively small effects on the hydrogen-bonded crystals, but it reduces the large MP2 errors for the imidazole and benzene crystals from ~10–12 to ~2–3 kJ/mol, which is only slightly worse than the CCSD(T) results; see [103] for further details.

Table 1 Molecular crystal lattice energies in kJ/mol. MP2C/CBS-limit performs much better than MP2/CBS-limit relative to both the estimated complete-basis-set CCSD(T) results and experiment

3.1.2 Crystal Structures

The HMBI model also provides accurate crystal geometries. Fully relaxed geometry optimizations at the HMBI counterpoise-corrected RI-MP2/aug-cc-pVDZ and Amoeba MM level for several of the benchmark crystals described above reproduced the experimental structures fairly accurately, including the space group symmetry [73]. Root-mean-square errors in the unit cell lattice parameters are only 1.6% relative to low-temperature crystal structures (100 K or below). For comparison, the same errors with B3LYP-D* [118] are 3.4% in the 6-31G** basis and 2.0% in the TZP basis. The HMBI RI-MP2 calculations also perform well for the root-mean-square deviations in the atomic positions (Table 2). Figure 8 shows overlays of the experimental and HMBI RI-MP2 optimized structures, highlighting the good agreement between them.

Table 2 Root-mean-square deviations (rmsd15 [140] in Å) between the theoretically optimized and experimental geometries for clusters of 15 molecules taken from each crystal. The temperatures at which the experimental structures was obtained is also listed. See [73] for details

Overall, the RI-MP2/aug-cc-pVDZ structures are clearly better than dispersion-corrected B3LYP-D*/6-31G* [118]. They are slightly worse than those obtained with B3LYP-D* in a triple-zeta basis (TZP), though the MP2/aug-cc-pVDZ results exhibit slightly more uniform errors. In principle, larger-basis MP2 geometries ought to be even better, but optimizing in those larger basis sets becomes even more expensive. One must also worry about the effects of the MP2 treatment of dispersion on the geometries.

Given the sorts of practical calculations that are currently feasible, it is not obvious that fragment methods provide a significant advantage relative to DFT for molecular crystal structure optimization, especially since the basis set requirements for correlated wavefunction methods like MP2 are steeper than those for DFT. For this reason, we often optimize the geometry using DFT and perform high-level single-point energies using HMBI. Of course, further improvements in MP2-like algorithms and continuing advances in computer hardware may change this.

Fig. 8
figure 8

Overlays of the experimental (black) and HMBI MP2 structures (red) for (clockwise from top left) ice Ih, formamide, acetamide, imidazole, and benzene. The experimental unit cell boundaries are drawn. Adapted with permission from [73]. Copyright 2012, American Institute of Physics

3.2 Aspirin Polymorphism

The crystal structure of aspirin has been known since the 1960s [119], but suspicion of a second crystal form lingered for many years until 2004, when the structure of aspirin form II was predicted by the Price group [120]. Experimental confirmation for form II came a year later [121], though this report was followed by several more years of controversy in the crystallographic community [122, 123]. Only in the past couple years has the existence and structure of form II aspirin been firmly established [124126].

The controversy arose from the fact that the two aspirin polymorphs exhibit very similar structures, and it proved very difficult to obtain pure crystals of form II. Rather, one often obtains mixed crystals containing domains of both forms I and II. The overall crystal packing of the two polymorphs is very similar except for the nature of the interlayer hydrogen bonding (Fig. 9). Whereas form I exhibits dimers, with each acetyl group hydrogen bonding to one other aspirin molecule in the adjacent layer, form II exhibits a catemeric structure, where each acetyl group hydrogen bonds to two adjacent molecules. This catemeric structure produces long chains of hydrogen bonds in form II.

Fig. 9
figure 9

Aspirin forms I and II exhibit very similar crystal packing. The key difference lies in whether the interlayer hydrogen bonding occurs as dimers (form I) or catemers (form II). Reprinted with permission from [132]. Copyright 2012 American Chemical Society

Interestingly, earlier dispersion-corrected DFT studies suggested that form II was ~2–2.5 kJ/mol more stable than form I [127, 128], which would be surprising for two crystals that appear to grow together so readily. Other examples of crystals which form intergrowths are often separated by less than 1 kJ/mol in energy [129131].

We investigated this system [132] using HMBI after optimizing the crystal structures for each form using B3LYP-D*/TZP. We performed a variety of calculations, including MP2, SCS(MI)-MP2 [104], and MP2C in both aug-cc-pVDZ and aug-cc-pVTZ basis sets. In all cases, the energy differences between the two polymorphs were only a few tenths of a kJ/mol, confirming the expected near-degeneracy of the two forms (Table 3). Interestingly, both MP2 and SCS(MI)-MP2 appear to overestimate significantly the ~115 kJ/mol experimental lattice energy [128, 133], predicting aug-cc-pVTZ values of 132 and 136 kJ/mol, respectively. In contrast, MP2C predicts a lattice energy of 116 kJ/mol, in much better agreement with the experimental value [103].

Table 3 Lattice energies and relative energy differences for the two polymorphs of aspirin (in kJ/mol). The two polymorphs are virtually degenerate across a range of model chemistries

In addition to predicting the virtual degeneracy of the two aspirin polymorphs, physical insight into the nature of the polymorphism was obtained by decomposing the relative energies of the two polymorphs according to the different contributions in the HMBI many-body expansion. In particular, we found that while the intramolecular interactions favor form I, the intermolecular interactions favor form II (Fig. 10). The key differences arise from the nature of the interlayer hydrogen bonding. In form I, the acetyl group hydrogen bonds to only one adjacent aspirin molecule, which allows the acetyl group to adopt a slightly more favorable intramolecular conformation. In contrast, the catemeric hydrogen bonding in form II forces the acetyl group to adopt a conformation that is slightly less favorable, but in doing so it forms much better hydrogen bonds and achieves hydrogen bond cooperativity through the extended hydrogen-bond networks. It turns out that the energy differences in each case amount to about 1.5 kJ/mol and cancel one another almost perfectly, making the two polymorphs virtually and “accidentally” degenerate.

Fig. 10
figure 10

Aspirin provides a classic example of conformational polymorphism: form I adopts a slightly more favorable intramolecular conformation, while form II exhibits stronger intermolecular interactions. The two effects have nearly identical energies, and the two polymorphs are virtually degenerate. Reprinted with permission from [132]. Copyright 2012 American Chemical Society

From a theoretical perspective, one other key feature emerged from this study: the strong similarity between the crystal packing in both polymorphs leads to excellent cancellation of errors in predicting the relative polymorph energetics. Unfortunately, such thorough error cancellation is probably the exception rather than the rule, as demonstrated by the case of oxalyl dihydrazide which is discussed in the next section.

3.3 Oxalyl Dihydrazide Polymorphism

Oxalyl dihydrazide provides another interesting example of molecular crystal polymorphism. It exhibits five experimentally known polymorphs, denoted α – ϵ, that differ in their degree of intramolecular and intermolecular hydrogen bonding (Fig. 11). The relative stabilities of the five polymorphs is not known experimentally, but it is believed that the trend follows α, δ, ϵ < γ < β [134]. That is, the β form is the least stable, followed by the γ form. The other three are more stable, though their energetic ordering is unknown.

Fig. 11
figure 11

The five experimentally known polymorphs of oxalyl dihydrazide

Earlier force field and DFT calculations had significant trouble reproducing these qualitative trends. The relative polymorph ordering varied widely with the choice of density functional; see, for example, [135]. Most of those functionals lacked van der Waals dispersion corrections, however. The empirical dispersion-corrected D-PW91 functional, which has been very successful in the blind tests of crystal structure prediction and elsewhere [12, 15, 87, 136138], does however obtain a plausible ordering for the polymorphs (α < ϵ < δ < γ < β), but the overall energy range of the polymorphs is ~15 kJ/mol, which is somewhat larger than the <10 kJ/mol typically found for experimentally observable polymorphs. On the other hand, a different dispersion corrected function, B3LYP-D* [118], gives a very different ordering that is inconsistent with experiment [70]. In other words, obtaining reasonable results for this system is not simply a matter of including dispersion. Rather, the energetics are very sensitive to the specific treatment of the electronic structure and the balance between intermolecular and intramolecular interactions. Note that vibrational zero-point energy is also important in oxalyl dihydrazide, and it is included in the results shown here. Finite-temperature effects may also be significant, but no attempt has been made to estimate them.

Obtaining plausible predictions for the relative polymorph energies proved challenging, even for correlated wavefunction methods (Fig. 12) [70]. In small basis sets, MP2 predicts the α form to be the least stable. However, increasing the basis set size toward the CBS limit dramatically reorders the polymorphs, preferentially stabilizing the α form. The slow basis set convergence in oxalyl dihydrazide provides a sharp contrast to aspirin, where the relative energy difference between the two polymorphs was insensitive to the basis set.

Fig. 12
figure 12

Relative polymorph energies for oxalyl dihydrazide as a function of the model chemistry. Basis set effects, improved MP2C two-body dispersion, and AIFF three-body dispersion are all important. Adapted with permission from [70]. Copyright 2012 American Chemical Society

The basis set sensitivity in oxalyl dihydrazide arises from BSSE. The α polymorph exhibits only intermolecular hydrogen bonds, while the other four forms contain mixtures of intermolecular and intramolecular hydrogen bonds. Applying a standard counterpoise correction to each QM dimer calculation helps correct the intermolecular BSSE, but it is harder to correct intramolecular BSSE. Therefore, the α polymorph, which has much less intramolecular BSSE, is destabilized relative to the other four forms in small basis sets. As the basis set size increases, the intramolecular BSSE decreases in the other four forms, and the balance between intermolecular and intramolecular hydrogen bonding is restored.

Beyond the basis set effects, oxalyl dihydrazide exhibits some π-stacking type van der Waals dispersion interactions, so it comes as no surprise that using MP2C instead of MP2 also has a significant effect. The dispersion correction reorders the β and γ polymorphs to the correct experimental ordering. It also stabilizes the α form substantially relative to the other polymorphs. The more empirical SCS(MI)-MP2 did not perform as well as MP2C in this system [70], perhaps due to the major differences in the optimal correlation energy scaling factors for intra- and intermolecular interactions.

Finally, Axilrod–Teller three-body dispersion effects prove important here. They destabilize the α form relative to the other four forms. The three-body dispersion contributions here are repulsive and die off with R − 9. Therefore, they are more repulsive in the dense α polymorph (1.76 g/cm3) than in the four other, less dense polymorphs (1.59–1.66 g/cm3). Without three-body dispersion, the calculations predict that the α polymorph is the most stable. However, including the three-body dispersion terms makes the ϵ polymorph slightly more stable than the α one, in contrast with the best dispersion-corrected DFT predictions. The predicted energetic preference for the ϵ polymorph over the α one is slight, and no experimental data currently exists to help resolve the issue. Nevertheless, it is interesting to note that the empirically dispersion-corrected DFT results in [70, 135] include only two-body dispersion and favor the α polymorph, just like the HMBI results without three-body dispersion. From this perspective it would be interesting to see what state-of-the-art many-body dispersion-corrected density functionals predict here.

4 Conclusions and Outlook

Fragment-based methods like HMBI provide a computationally viable means of achieving high-accuracy structures and lattice energies for chemically interesting molecular crystals. The systems examined here highlight the challenges inherent to modeling molecular crystals, where the subtle energetic competitions can require careful electronic structure treatments. It will not always be a priori obvious how elaborate an electronic structure treatment is needed for a given system. One can sometimes count on cancellation of errors, especially when the different crystal packing motifs are similar (like in aspirin), but much less error cancellation occurs in other cases (like oxalyl dihydrazide). Unfortunately, the reduced error-cancellation case is probably more typical for conformational polymorphs of flexible organic molecules, especially as the molecules become larger.

Periodic DFT methods can often provide acceptable accuracy for molecular crystal systems with reasonable computational costs. In many cases, however, one will wish to assess the reliability of those predictions or to obtain even more accurate predictions. From this perspective, one of the key strengths of fragment methods is their ability to improve systematically the quality of the results with respect to the underlying electronic structure treatments. The ability to demonstrate convergence of the predictions with respect to model chemistry is a powerful tool for making robust predictions. Moreover, for systems with large numbers of molecules in the unit cell, the inherently linear-scaling nature of the computationally dominant QM calculations in fragment methods will often make them cheaper than full periodic DFT calculations. Fragment methods also provide a natural decomposition scheme for understanding the nature of the energetic competitions in polymorphic crystals.

For evaluating the intermolecular interactions in molecular crystals, the dispersion-corrected MP2C method provides a useful balance between accuracy and efficiency, offering near coupled-cluster-quality results at MP2-like cost. In practice, MP2C (or its explicitly correlated variant, MP2C-F12), with an aug-cc-pVTZ basis, is practical for molecular crystals containing two to three dozen atoms per molecule and a handful of molecules in the asymmetric unit cell. In other words, these techniques are applicable to a number of interesting, small-molecule pharmaceuticals, organic semiconductors, and energetic materials (though many more such materials remain impractical for the moment!).

Thus, fragment methods like HMBI will likely play an important role in molecular crystal modeling for years to come. The next advances are likely to come from a couple of directions. First, further efficiency improvements will enable the application to larger systems. For instance, the MP2C timings described above indicate that the vast majority of the computational time is consumed on the MP2 calculations. A sizable fraction of that time comes from evaluating the long-range interactions, which are then largely discarded through the MP2C dispersion correction. New, more efficient strategies for achieving similar quality results can surely be developed. In addition, further improvements to the ab initio force field would enable one to treat fewer dimers quantum mechanically, thereby accelerating the calculations. This requires incorporating efficient treatments of the short-rate exchange interactions into the force field.

Second, one needs to consider finite-temperature entropies and free energies. This can be done through quasi-harmonic-type approximations or through dynamical free energy calculations. The latter is potentially more rigorous and capable of capturing anharmonic effects, but of course it is limited by the need for extensive computational sampling and the quality of the force fields that can be used to perform such sampling affordably. Quasiharmonic calculations are much less expensive computationally, but they involve their own severe approximations. Much more research is needed in this area to determine which techniques are useful under which circumstances. In other words, many theoretical opportunities remain in molecular crystal modeling to occupy researchers for quite some time!