Heterotic Orbifold Models
1 Universidad Nacional Autónoma de México, POB 20-364, Cd.Mx. 01000, México 2 Department of Physics and Astronomy, University of California, Irvine, CA 92697-4575, USA |
Invited chapter for the Handbook of Quantum Gravity (edited by Cosimo Bambi, Leonardo Modesto, and Ilya Shapiro, Springer 2023) |
Abstract
We review efforts in string model building, focusing on the heterotic orbifold compactifications. We survey how one can, starting from an explicit string theory, obtain models which resemble Nature. These models exhibit the standard model gauge group, three generations of standard model matter and an appropriate Higgs sector. Unlike many unified models, these models do not suffer from problems such as doublet-triplet splitting, too rapid proton decay and the problem. Realistic patterns of fermion masses emerge, which are partly explained by flavor symmetries, including their modular variants. We comment on challenges and open questions.
Keywords
String compactifications, Heterotic string, model building, phenomenology, symmetries of string models.
1 Introduction
If string theory is to describe the real world, it has to reproduce our current established understanding of physics. In particular, its low-energy description has to give rise to the standard model (SM). Generally, string model building concerns the question of how the SM fits into string theory. In practice, one compactifies a consistent string theory to a four-dimensional model which can be studied and confronted with observation. One particularly important aspect of top-down model building is that the globally consistent models are complete, i.e. they do not only describe the SM but also include, say, the degree(s) of freedom driving cosmic inflation and dark matter. That is, unlike in the bottom-up approach, one cannot add extra sectors to the model at will.
Historically, the first attempts to construct realistic string models were based on the heterotic string. It was noticed that the structure of SM is remarkably consistent with unification along the exceptional chain [4]. In this review we provide a brief overview of orbifold compactifications of the heterotic string which come close to the SM.
Heterotic models come broadly in two classes, they can either be based on smooth compactifications [5] or on obifolds [6, 7], cf. Figure 1. These classes are related as some smooth compactifications can emerge from orbifolds via blow-up (cf. e.g. [8]). Orbifolds have the advantage that their construction involves explicit strings, which is why they will be our focus. Orbifolds can be constructed in the so-called free fermionic approach, yet our focus will be the classical approach, in which the so-called symmetric orbifolds have a geometric interpretation.

The purpose of this review is to summarize the current status of heterotic model building. There are already excellent reviews of this subject such as [9], however, our focus will be on more recent developments and a clear account of the open questions. To this end, we will review the target of string model building, the SM and some of its extensions in Section 2 before turning to the heterotic string and its compactifications in Section 3. In Sections 4 and 5 we collect some facts about the spectra and symmetries of the constructions, which will be the basis for the discussion of challenges in Section 6. In Section 7 we provide some explicit examples. After briefly commenting on smooth heterotic compactifications in Section 8, we provide an outlook in Section 9.
2 What do we (believe to) know?
Before delving into what string theory gives us, let us briefly survey what we expect to get out of string model building.
2.1 A very short recap of the SM
First and foremost, we wish to obtain a quantum field theory (QFT) that is consistent with the SM (see e.g. [10] for a detailed description). The latter is based on the continuous gauge symmetry111Strictly speaking we do not really know the gauge group of the SM but only its Lie algebra, a subtlety which we will, like most of the literature, not discuss in detail.
(1) |
The matter content consists of three generations of quarks and leptons, left-chiral Weyl fermions which transform as
quarks | (2a) | |||
leptons | (2b) |
under . Here, labels the generations. In addition, there is the so-called Higgs field, a complex scalar carrying the quantum numbers . The Higgs acquires a vacuum expectation value (VEV), , which breaks down to , under which (2) are vector-like and acquire masses which are given by the product of so-called Yukawa couplings and . In the SM, the Yukawa couplings are input parameters which are adjusted to fit data. A curious fact about the SM is that the combination of charge conjugation and parity, , is broken in the flavor sector, i.e. by the Yukawa couplings, but seemingly not in the strong interactions. This mismatch gets referred to as the strong problem. The neutrinos, which are part of the , are also massive, yet it is currently not known which operator describes their mass. The most plausible options are the Weinberg operator, , or Dirac neutrino masses, in which case one has to amend (2) by right-handed neutrinos . The neutrino masses are much smaller than the masses of the charged fermions.
It is instructive to survey the continuous parameters of the SM. They comprise (i) gauge couplings; (ii) ; (iii) Higgs parameters; (iv) masses; (v) or mixing parameters, depending on whether neutrinos are Dirac or Majorana particles. In a stringy completion, these parameters should be predicted rather than adjusted, and, as we shall discuss in Section 6, the requirement to reproduce these observables remains one of the greatest challenges in string model building. Note also that currently the only other parameters that we need to describe observation are the Planck mass (or, equivalently, ), the vacuum energy and the density to dark matter, . That is, currently the bulk of the (ununderstood) parameters of Nature resides in the flavor sector of the SM.
An important fact about the SM is that it does not only provide us with couplings and interactions that have been confirmed in experiments, but it also comes with tight constraints on additional particles and interactions. In particular, it is extremely hard to make extra states which are chiral w.r.t. consistent with observation. Also, while the Weinberg operator is a nonrenormalizable operator that is “good” in the sense that it can describe neutrino masses, other higher-dimensional operators are highly constrained. For instance, the suppression scale of the dimension-6 operators leading to proton decay has to exceed .
2.2 Early universe
The early universe provides us with important insights into high energy physics (see e.g. [11]). For instance, big bang nucleosynthesis (BBN) works very well within the SM, and extra particles may be inconsistent with the primordial formation of the elements if they decay late or increase the Hubble expansion rate too much. In addition, fractionally charged particles are often stable since they cannot decay into SM states, and there are stringent constraints on their relic abundance (cf. e.g. [12, 13]).
However, the early universe also requires physics beyond the SM. Most notably we need a field or sector that drive inflation, or another ingredient which provides us with solutions to the so-called horizon and flatness problems. In addition, there is very compelling evidence for dark matter which cannot be made of SM particles. Furthermore, the baryon asymmetry of the universe requires physics beyond the SM, too.
2.3 \AcBSM scenarios
Having seen that physics beyond the SM is required to accommodate astrophysics and cosmology, let us spend some words on beyond the standard model (BSM) scenarios.
The supersymmetric variants of the SM (see e.g. [14] for an introduction), most notably the minimal supersymmetric standard model (MSSM), have received substantial attention in the past decades. This is because, assuming low-energy supersymmetry (SUSY), the electroweak scale gets stabilized against quantum corrections. Of course, given the absence of clear signals for SUSY at the Large Hadron Collider (LHC), this scheme has lost some of its popularity in the recent years, yet the MSSM is arguably still one of the best motivated and well-defined BSM scenarios. The MSSM has a number of shortcomings which one may hope to solve in ultraviolet (UV) completions, and the purpose of this review is to discuss these solutions in Section 7. In order to understand some of these shortcomings, let us look at the invariant superpotential terms up to order 4,
(3) |
where we, in a slight abuse of notation, denoted the superfields by the same symbols as the SM fields in Equation 2. Note also that the MSSM has two Higgs doublets, and . The couplings , and are the Yukawa couplings which yield the masses of quarks and charged leptons. The -parity violating terms , , and have to be highly suppressed, and get often forbidden by (or matter) parity, the origin of which is to be clarified in a UV completion of the model. The term in the first line of Equation 3 can a priori have any size, but in order to have proper electroweak symmetry breaking and sufficiently heavy Higgsinos it should be of the order TeV or so, which is a common choice for the soft SUSY breaking masses. Explanations of this fact comprise the Kim–Nilles [15] and Giudice–Masiero [16] mechanisms, and we will see later in Section 7 both are realized in explicit stringy completions of the MSSM. Further, while the term in the last line of Equation 3 can describe neutrino masses, the terms have to be very small, e.g. the coefficients and have to be suppressed by more than , where . A proper understanding of this suppression arguably requires a solution within a consistent theory of quantum gravity, such as the explicit string models we consider here.
Another appealing feature of the MSSM with low-energy SUSY is that gauge couplings unify remarkably well at a scale [17]. This has led to the scheme of SUSY Grand Unified Theorys (see e.g. [18] for an extended discussion), in which a unified symmetry like or gets broken at down to . An arguably even stronger motivation for GUTs is the structure of matter since one generation of the SM (cf. Equation 2) including the right-handed neutrino its into a -plet of ,
(4) |
where we also indicated the representations in underbraces. While the GUT symmetries work very well for the matter, they fail for the SM Higgs. The smallest representations that contain the MSSM Higgs doublets are , which combine to a -plet of . The additional -plets contained in these representations pose major threats to the model as they typically mediate unacceptably large proton decay unless their mass exceeds the Planck scale. This is one facet of the doublet-triplet splitting problems which haunt GUTs in 4D. On the other hand, as has been pointed out early on in the context of string model building, in higher-dimensional, in particular stringy, models the same mechanism that breaks the GUT symmetry can also split the doublets from the triplets [5, 19].
2.4 What do we hope to learn from string model building?
Of course, it will be reassuring to find a compactification which reproduces the SM in great detail, regardless of whether or not the underlying construction is unique. What is more, given such a model, we will be in a unique position to answer some of the most popular questions of our time:
-
1.
What is the origin of flavor and violation?
-
2.
What is the nature of dark matter and what are the properties of the dark (aka hidden) sector?
-
3.
What drives cosmic inflation?
While these questions might be answered separately, the power of addressing them in explicit string models is that the answers are much more specific and related in intriguing ways.
3 Compactifying the heterotic string
3.1 Heterotic string
In this section, we collect some basic facts on the heterotic string. For further details and a broader overview see [20]. The term ‘heterotic’ derives from the Greek word ‘hetero’, which translates as ‘other’, and in biology is related to ‘vigorous hybrid’, which arguably reflects the nature of the heterotic string. The heterotic string theory [21] is the result of combining a 10D superstring and a 26D bosonic string. The former can equip the theory with supersymmetry in ten dimensions whereas the bosonic string provides us with a non-Abelian gauge group of rank 16,222The nonsupersymmetric heterotic string can be obtained from this version, as we briefly describe in Section 7.2.
(5) |
Note that the most general compactification can have continuous enhancements of these gauge symmetries, yet we will mainly focus on (5). The heterotic theories contain only oriented closed strings propagating in ten dimensions.
In lightcone gauge, there are 8 right-moving bosonic string coordinates and 8 right-moving fermions , where . and denote the worldsheet coordinates. There are in total 24 left-moving coordinates . In symmetric compactifications they get decomposed into with as in the right-handed sector, and with . This decomposition gives rise 8 combinations of ordinary physical coordinates, . The additional left-moving coordinates are responsible for the gauge symmetries , cf. Equation 5.
For the sake of keeping this review short, we specialize on symmetric compactifications, cf. Figure 1. Interestingly, the so-called Free Fermionic Formulation (FFF) and geometric orbifolds are related by a dictionary [22, 23]. There are possibilities to go more general, and consider e.g. asymmetric orbifolds [24, 25] or Gepner models [26, 27], which is beyond the scope of this review.
3.2 Heterotic strings on orbifolds
We will start our discussion with heterotic orbifolds [6, 7], which allow one to explicitly “see” the strings. For simplicity, we focus on symmetric toroidal orbifolds, which emerge by dividing tori by some of their symmetries. The tori are given by or , where denotes a lattice, or, more precisely, the group of lattice translations. We will be interested in . Therefore, with denoting the so-called space group, which is comprised of lattice translations and additional operations such as rotations and so-called roto-translations, and forms a discrete subgroup of the -dimensional Euclidean group. Crucially, these operations are also embedded in the gauge sector, which breaks down to a subgroup. Moreover, they also break SUSY, which facilitates the construction of chiral 4D models with or no SUSY.
In more detail, in the geometric formulation elements of space group are conveniently denoted by , where . The () are the basis vectors of the underlying torus. The set of form a finite group, called the point group , which determines the holonomy group, and thus the amount of SUSY that survives the compactification. In fact, in order to classify the physically inequivalent orbifolds, one only needs to find the different affine classes [28], but we refrain from spelling this discussion out in detail. If is Abelian, it is either or . If in addition SUSY is preserved, then a given transformation can be encoded in a so-called 3-component twist vector , which describes the rotations of three complex coordinates. In general, acts on string coordinates as
(6) |
The space group is to be embedded in the gauge degrees of freedom. Loosely speaking, the point group elements get mapped to so-called shift vectors . This embedding has to preserve the order, i.e. if satisfies then , where denotes the root lattice of , cf. Equation 5. In orbifolds, if, say, the shifts of two models differ by lattice vectors , the resulting models are identical (cf. [29]). This is no longer true in orbifolds, where gauge embeddings differing by lattice vectors may be inequivalent [30], and be related via what is known as discrete torsion [31]. Since is even and self-dual, one can find an Euclidean basis in which the lattice vectors are given by
(7) |
where , and denotes the dimensions of the Lie algebras or , respectively. The gauge embedding of each translation is a so-called discrete Wilson line . The Wilson lines are constrained by geometry. In more detail, since lattice vectors get mapped onto each other by the rotations, the analogous relations have to hold for the Wilson lines,
(8) |
For instance, in a orbifold plane one has and . Therefore, and , where “” means “equal up to ”. Thus . This generalizes to other geometries, i.e. for a given Wilson line there is an integer such that , with no summation over . As a consequence, the coefficients in (8) are integer.
In addition, the orbifold parameters and their gauge embeddings must satisfy a series of constraints in order to ensure world-sheet modular invariance, which guarantees the UV consistency of the model [7, 31]. For orbifolds, these conditions take the form [30]
(9) |
where no summation over nor is implied.
3.3 Classification of toroidal orbifold geometries
While early classifications of viable toroidal orbifolds focused on special kinds of lattices [32], more recently a richer set of possibilities has been uncovered in the orbifold [22] and generalized to other point groups [28, 33]. Loosely speaking, the new ingredient of the additional possibilities are space groups which contain roto-translations but (cf. [34]). As a consequence, the fundamental group of the orbifolds (and not just the underlying tori) can be nontrivial. Among other things, this allows for non-local, or Wilson line, breaking of the gauge symmetry, which is also being utilized in the context of smooth compactifications [35]. Another innovation is the consistent construction of non-Abelian orbifolds [36, 33]. In particular, there are 138 Abelian and 331 non-Abelian space groups [28] of toroidal symmetric orbifolds preserving SUSY in 4D. These geometries have been shown to host many models with gauge symmetry and chiral spectrum of the MSSM [37, 38, 39], yet their detailed phenomenological properties have not been worked out so far.
3.4 Anisotropic compactifications
Because of the gauge symmetries of the heterotic string, heterotic models comply well with the idea of grand unification. Breaking the GUT symmetry via compactification allows one to elegantly avoid the major shortcomings of 4D GUTs, most notably the doublet-triplet splitting challenge and its associated proton decay problems. However, there is a tension between the scale of gauge coupling unification in the MSSM, , and typical compactification radii. This is because string theory also describes gravity, and the effective 4D Planck mass is sensitive to the volume of compact space. In some more detail, Newton’s constant is related to the fine structure ‘constant’ at the GUT/compactification scale, , and the string tension, , via [40]
(10) |
for a weakly coupled theory. This value of is too large for typical values of and (cf. our ealier discussion around Equation 4). There are various proposals to fix this issue (see e.g. [41]). The perhaps most ingenious way to address this problem is M-theory [40]. However, the problem can also be ameliorated in anisotropic compactifications [40, Footnote 3]. A detailed analysis [34] suggests that this solution barely fails, but by the own admission of the authors the presented bound is too conservative. In fact, if one uses the appropriate volume of the orbifold for the analysis rather than the underlying torus, one finds that anisotropic compactifications can work, even though the parameter space of solutions is not too generous. This implies that there is an intermediate orbifold GUT symmetry (see e.g. [42] for a review). However, this also means that the smaller radii are of the order of the string scale, and as stressed in [40], one must use conformal field theory (CFT) (rather than classical geometry) to analyze the model. This is one of the reasons why this review focuses on orbifold constructions.
4 Spectrum
Given a compactification of the heterotic string, one can determine its spectrum, i.e. the properties of the massless and massive excitations. One usually proceeds in two steps, by first determining the spectrum “after compactification” and then the spectrum of deformation of the model in which certain VEVs get switched on. In this section we focus on the former.
4.1 Massless gauge fields
In general, only a subset of the gauge fields survive the orbifold projections. They can be determined by finding the roots , i.e. , which satisfy the projection conditions
(11) |
with the from Equation 7 for all shift and Wilson line vectors (cf. Section 3.2).
4.2 Chiral zero modes
The zero modes are solutions of the mass equation to vanishing mass. In all known examples they are chiral w.r.t. some, possibly discrete, symmetry. The computation of the massless spectrum, i.e. gauge and chiral zero modes, is straightforward though tedious if done by hand, and can conveniently be performed with dedicated tools such as the Orbifolder [43].

Explicit string models exhibit states beyond the SM or MSSM at some level. The additional states include the moduli as well as the winding and Kaluza–Klein (KK) modes, which we will review in Sections 4.3 and 4.4, respectively. In addition, there are often vector-like states w.r.t. the SM gauge group which are neither KK nor winding modes. Whether or not these vector-like states are massless often depends on the point of moduli space under consideration. For instance, vector-like states may attain masses when giving VEVs to blow-up modes that smoothen out orbifold singularities, and break symmetries w.r.t. which these states are chiral. However, it would be arguably wrong to refer to the smoothened out version as “cleaner” since it is really the same model. In fact, often important properties of a given construction are much more directly accessible by studying the symmetry-enhanced point in field space, which is given by the orbifold point in this example, even though the vacuum is away from this point.
4.3 Moduli
Virtually every supersymmetric string compactification contains fields which are classically flat directions, and this is in particular true for models that come close to the real world. Some of these so-called moduli do not have any charge under , and comprise the Kähler moduli , the complex structure moduli and the dilaton . Yet also some of the charged fields can attain VEVs because they are along -flat directions. The VEVs of these fields determine, among other things, geometric properties of compact space. Classical flat directions can attain a nontrivial potential at the quantum level, in particular through nonperturbative effects. It is generally challenging to compute these potentials in full detail and thus determine the VEVs at the minima, see [44] for more details of the analogous discussion in other versions of string theory.
4.4 Winding and KK modes
5 Symmetries of the effective action
Heterotic compactifications lead to effective 4D theories exhibiting various symmetries [45], which largely determine the phenomenological properties of the respective models. These symmetries include
-
1.
either or no supersymmetry,
-
2.
continuous gauge symmetries, which mainly originate from the 10D symmetries, i.e. the root lattice of the 16 left-moving coordinates,
-
3.
symmetries (in SUSY compactifications),
- 4.
-
5.
modular symmetries, and
-
6.
outer automorphisms which may be or -like transformations.333It is important to distinguish between proper transformations, which map all particles to their own antiparticles, and -like transformations, which only send some of the particles to their antiparticles [46].
More recently, it has been pointed out that the symmetries 3-3 may be regarded as outer automorphisms of the Narain lattice [47, 48] in the Narain formulation of toroidal compactifications of the heterotic strings [49] (see also [50]). The gauge symmetries can go beyond if the compact space has special properties, such as some radii equalling certain critical values (cf. e.g. [50]).
5.1 SUSY
SUSY (cf. Section 2.3) has long been a standard ingredient of string models. Whether or not SUSY is preserved by the compactifaction depends on the holonomy group of the compact space [5]. In the case of a smooth compactification, the requirement that the compactification preserves SUSY dictates that the manifold has to be of the Calabi–Yau (CY) type, and in orbifolds it requires the twist to fit into , the holonomy group of CY manifolds.
5.2 Continuous gauge symmetries
After compactification the residual continuous gauge symmetry, , of a realistic model has to contain the gauge symmetry of the SM (1). The gauge symmetry follows already from our discussion in Section 4.1. Apart from the obvious option that promising models may also replace by the Pati–Salam [1] (PS) group , the so-called left-right symmetry [51] , or the flipped symmetry [52]. Other grand unified symmetries are in principle possible but may be challenged by doublet-triplet splitting problems and the lack of appropriate Higgs fields that break the larger symmetry to .
Very often in geometric orbifolds with SUSY one factor appears anomalous, with the anomaly being cancelled by the Green–Schwarz [3] (GS) mechanism. As a consequence, the -term potential contains an Fayet–Iliopoulos [2] (FI) term, , where
(12) |
denotes the gauge coupling, and the generator of . The requirement of a vanishing of the -term potential induces VEVs of charged fields that breaks and in the overwhelming majority of cases further symmetries [53]. Clearly, in realistic models the fields acquiring large VEVs must be SM singlets. Configurations with vanishing -terms can be identified by constructing holomorphic monomials which are invariant under all gauge symmetries but carry nontrivial charge under [54, 55, 56]. A complete basis of such monomials can be obtained via the Hilbert basis [57]. In the vast majority of explicit models, , so that is of the order of the Cabibbo angle, and, therefore, the VEVs induced by (12) may conceivably play a role in explaining flavor hierarchies [58].
The extra symmetry is usually partly broken by the VEVs that cancel the FI term. The residual continuous part can conceivably provide us with a hidden sector leading to dynamical SUSY breakdown [59, 60]. There are also usually discrete symmetries, which can be determined systematically with the Smith normal form [61].
5.3 Discrete symmetries
5.3.1 Symmetries from a Narain compactification
The Narain formulation provides an alternative to the usual toroidal compactification of the heterotic string discussed in Section 3. Let us consider first scenarios in which the six extra dimensions are compacitified in a . In the Narain formulation, the right- and left-moving compact (bosonic) coordinates are considered independently, so that, taking into account also the gauge degrees of freedom, the compactification is specified in terms of an auxiliary D torus according to
(13) |
Here, is the Narain vielbein which spans the D even, integer and self-dual Narain lattice of signature . Further, the integer vector includes the winding numbers , the KK numbers , and the gauge momenta discussed in Section 5.2. The properties of are encoded in the condition
(14) |
where is the flat metric with signature , denotes the 16D Cartan matrix of the heterotic string, and we have defined the Narain metric .
It turns out that the group of outer automorphisms of , defined by
(15) |
describes all the discrete symmetries of the toroidal compactification. Hence, naturally contains the modular transformations of the compactification, including mirror symmetries and -like transformations of the moduli.
From this compactification, it is easy to arrive at the symmetries of a toroidal orbifold. In this formalism, an orbifold is obtained by modding out a subgroup of . Let us consider, for simplicity, the case of an Abelian orbifold without roto-translations. Treating left- and right-moving coordinates as independent, as before, the orbifold identification is given by
(16) |
with the Narain twist
(17) |
We can further impose the orbifold to be of order by demanding . The Narain twist must leave the chosen Narain lattice invariant, i.e. , which ensures that the moduli remain invariant under the orbifold action. Hence, some of the moduli of the original toroidal compactification are hereby fixed. Note that the possibility defines an asymmetric orbifold. Limiting ourselves to , we recover the geometric picture of the symmetric orbifolds introduced in Section 3.
The discrete symmetries of the orbifold include then the subgroup of rotational outer automorphisms of the toroidal compactification, , that are left unbroken by the orbifold, i.e. which satisfy
(18) |
and is the Narain twist in the Narain lattice basis. In addition, now there are translational outer automorphisms of the orbifold444The Narain twist combines with the translations of the Narain lattice to build the Narain space group . Formally, it is the outer automorphisms of that we refer here as the automorphisms of the orbifold. given by
(19) |
In order to be compatible with the orbifold, the translations must fulfill
(20) |
Note that these translations build a normal subgroup of the full group of outer automorphism of the orbifold.
These discrete residual transformations give rise to , flavor, modular and outer automorphism symmetries, which we will discuss separately in what follows.
5.3.2 symmetries
Supersymmetric orbifold compactifications usually do not break the Lorentz symmetry of the compact 6 dimensions completely but leave discrete remnants which act as symmetries in the effective description. Since the superpotential has a nontrivial modular weight, modular transformations, which we will discuss in more detail below in Section 5.3.4, are generically symmetries. As we shall see in an explicit example in Section 7.1, certain symmetries can be instrumental in resolving some of the phenomenological issues.
5.3.3 Flavor symmetries
The repetition of families in the SM begs for an explanation. Flavor symmetries may address this question. String compactifications can give rise to non-Abelian discrete symmetries in which the three generations of the SM transform as a -plet, or two generations as a -plet. Such symmetries may arguably play a role in understanding the flavor structure of the SM (see e.g. [62] for references).
In the geometric approach, flavor symmetries can be obtained from the replication of matter states at different yet equivalent orbifold singularities in the compact dimensions [30]. The emerging permutations combine with additional symmetries from the string selection rules to non-Abelian discrete symmetries. These rules act on matter fields as Abelian symmetries of the effective theory, which can be understood as an Abelianization of the space group of the orbifold [63]. It has been verified in explicit examples that the above-mentioned non-Abelian symmetries emerge from continuous gauge symmetries [64] are hence gauged, as one would expect.
In the Narain formalism, these symmetries are identified with the subgroup of translational outer automorphisms of the orbifold.
5.3.4 Modular symmetries
Modular symmetries are ubiquitous in string compactifications. They are symmetries of certain loop diagrams and the partition function. Moreover, toroidal orbifold compactifications exhibit modular symmetries. It is important to distinguish between the two.
World-sheet modular invariance has far-reaching implications for the UV consistency of the theory, the comprehensive discussion of which is beyond the scope of this review. In particular, modular invariance conditions constrain the choices of the geometrical data of the models [7, 31]. Among other things, they ensure that the models are free of anomalies.
Target-space modular invariance provides us with important constraints on the Kähler potential and couplings of the theory [65]. These modular symmetries contain crucial information on the couplings of the theory [66], and even provide us with an alternative to the CFT computation [67].
In the geometric approach of toroidal orbifold compactifications, the properties of target-space modular symmetries have been explored. Among other features, these symmetries are free of anomalies thanks to the GS mechanism. Further, the transformation of matter fields under these symmetries have been determined. Denoting by555Here, we adopt the convention , where is the nontrivial component of the antisymmetric -field and the metric of the 2D orbifold sector, respectively. the Kähler modulus of a orbifold sector, by any transformation from the corresponding modular group, the th multiplet of twisted matter fields of the orbifold transform according to [68, 69]
(21) |
Here is a representation of in a finite (double cover) modular group with depending on the order of the orbifold, and is the (possibly fractional) modular weight carried by the twisted fields [70].
The 4D effective supersymmetric field theory of such a model is governed by these symmetries. In particular, the modular transformation of the associated Kähler potential reads at leading order
(22) |
This transformation is cancelled by a Kähler transformation of the superpotential provided that the superpotential terms of order are given by
(23) |
and have total modular weight per complex orbifold plane, where is a modular form.
In the Narain formalism, the modular symmetries are identified with the subgroup of rotational outer automorphisms of the orbifold. Note that the symmetries (cf. Section 5.3.2) may also be understood as remnants of the target-space modular transformations of the complex structure moduli of the orbifold. This implies a relation between the charges of matter fields and their so-called modular weights [71].
5.3.5 Outer automorphisms
The effective action can exhibit certain outer-automorphism symmetries. These symmetries contain fundamental transformations like charge conjugation , parity and time reversal . has to be broken in the flavor sector order to describe the real world, a criterion that already some of the first explicit string models turn out to satisfy [72]. Further outer automorphisms comprise the left-right parity of the left-right symmetric model, which may emerge as discrete remnants of the continuous gauge symmetries after orbifolding [73].
Note that in the Narain formalism some outer automorphisms of the orbifold can also be considered -like transformations. It remains to be seen whether there is a connection between such transformations and the physical .
5.4 Approximate symmetries and hierarchies
Starting at a symmetry-enhanced point has various benefits compared to analyzing generic points in moduli space. The models reviewed in this review give rise to a variety of mildly broken, and thus approximate, symmetries. As already mentioned above, the latter may conceivably explain the observed hierarchies in the flavor sector [58]. They may also provide us with solutions to the and/or strong problems (cf. [74, 75]). They may explain the scales in models of dynamical supersymmetry breaking (such as [60], which requires an explicit mass for some pairs of vector-like states) or the messengers of gauge mediated SUSY breaking [76].
6 Challenges
Modern days model building faces various challenges. Bottom-up models usually can accommodate observation but the shear abundance and flexibility of the emerging constructions make it appear unlikely that these activities alone will provide us with unique answers. This review focuses on top-down models, which come with their own challenges. They include:
-
C1.
Obtain the correct gauge symmetry.
-
C2.
Obtain the correct spectrum, i.e. the three generations of quarks and leptons without any other states which are chiral w.r.t. .
-
C3.
Avoid dangerous operators such as those leading to too fast proton decay or too large flavor changing neutral currents.
-
C4.
Provide a consistent cosmological history.
-
C5.
Reproduce the observed values of continuous parameters of the SM, i.e. the gauge and Yukawa couplings.
The first challenge, C1, has been mastered successfully in heterotic model building early on. Compactification breaks the gauge symmetry of the 10D heterotic string, and it is fairly straightforward to obtain the SM gauge symmetry, or a symmetry that can be broken to .
Obtaining the correct spectrum, i.e. C2, has been a bigger challenge since string models may yield the wrong number of generations or give rise to chiral exotics. Nonetheless, extensive scans accompanied with appropriate search strategies has enabled the community to identify a large number of compactifications that exhibit the chiral spectrum of the SM at low energies.666However, the number of chiral generations at low energies and at the compactification scale may be different [77], so some care needs to be taken not to prematurely discard models. Notice that this often involves the appearance of extra, vector-like states which acquire masses below the compactification scale. While there are some constraints on such states, e.g. from the requirement that the gauge couplings remain perturbative, they also may play an important role in the phenomenology of the model, cf. our discussion in Section 4.2. Another concern stems from so-called fractionally charged exotics. As already mentioned in Section 2.2, there are tight experimental constraints on their relic abundance (cf. e.g. [12, 13]), so the appearance of such states in the spectrum leads to the requirement that they are not produced copiously in the early universe.
The offending operators mentioned in C3 include the so-called -parity violating couplings of the MSSM, cf. our discussion below Equation 3. Forbidding this couplings requires additional symmetries, the simplest option being - or matter parity. This shows, in particular, that it is not sufficient to obtain models with three generations and the SM gauge symmetry, one necessarily needs additional symmetries. The offending operators also often get induced by extra states. For instance, triplet partners of the MSSM Higgs doublets may mediate proton decay at an unacceptably large rate [78]. This problem haunts 4D models of grand unification but is absent in certain higher-dimensional variants [79]. Nonetheless, it may get reintroduced through other vector-like states.
While string cosmology is an active field (see [44] for more details), only limited attention has been given to the performance of otherwise promising models, which overcome C1, C2, C3 and C3. It remains a task for the future to see whether, say, inflation can be realized and a realistic baryon asymmetry can be generated.
Challenge C5 is major, and has not been completely mastered in any known construction so far, let alone in remotely realistic models. Part of the problem is that the couplings depend on the VEVs of certain scalar fields, the moduli (cf. Section 4.3) and possibly other fields.777The precise definition of what a modulus is varies over the literature. In some parts -flat combinations of the other fields are referred to as moduli. This means that mastering challenge C2 requires stabilizing all moduli. This is a topic on its own, which is covered in [44]. Within the examples given in Section 7, we will comment on the extent to which realistic couplings are obtained.
7 Examples
7.1 Geometric orbifold with SUSY
Rather than reviewing extensive model scans, let us focus on a particular example, the model of [80]. The orbifold has noncontractible cycles (cf. Section 3.3) which allow one to break an grand unified symmetry nonlocally down to . This type of GUT breaking avoids fractionally charged exotics. The spectrum consists of 3 chiral generations of quark and leptons plus additional states which are vectorlike w.r.t. but massless at the orbifold point, see Table 1. Like the majority of models of this type, there is an FI term which has to be cancelled consistently with vanishing of the - and (other) -terms. The corresponding VEVs break the gauge symmetry at the orbifold point down to
(24) |
where and stem from two different factors, and none of the SM matter is charged under .
1 | 1 | 1 | 1 | 1 |
0 | 2 | 0 | 2 | 0 | 0 | 0 | 2 | 0 | 0 | 2 | 2 | 0 | 2 | 2 | 2 | 0 | 0 |
These VEVs also provide mass terms for all SM charged exotics, yet the symmetry forbids the mass of one linear combination of Higgs fields, which gets identified with the MSSM Higgs pair. This pair acquires a mass after breaking. The order parameter of symmetry breaking is the gravitino mass, i.e. of the order of the soft terms, which are assumed to be not too far above the electroweak scale. That is, the , which is a discrete remnant of the Lorentz symmetry of compact space, can provide us with a solution to the problem along the lines of [81]. In addition, this suppresses dimension-5 proton decay operators enough to be consistent with observation.
It has been checked that qualitatively realistic fermion masses arise, i.e. the Yukawa couplings have full rank, exhibit hierarchies and lead to nontrivial flavor mixing, and neutrino masses are see-saw suppressed. However, this is not to say that they are fully realistic, cf. our discussion of C5.
Altogether this example shows that explicit string models can successfully address some of the most pressing questions of (traditional) unified model building, including the and proton decay problems. However, it also illustrates that there is still a long way to go before we can claim to have found “the” stringy SM. Apart from the question whether or not low-energy SUSY is realized in Nature, one has to successfully fix the moduli. While this is a topic on its own, which is covered in [44], the symmetry and charges can be used to show that generically there are no flat directions. States with odd acquire masses because the mass terms carry charge , and the fields of charge 2 pair up with linear combinations of charge 0 fields. Of course, generic statements do not always lead to the correct conclusions, and one has to verify explicitly that there are no flat directions, what the possible VEVs of the charge 0 fields are, and whether they leads to phenomenologically viable couplings in the SM sector.
7.2 Geometric orbifolds without SUSY
There is a consistent nonsupersymmetric heterotic string [82, 7, 83], which can be understood as a freely acting orbifold of a heterotic string [7, 83]. Given the absence of evidence of supersymmetry at colliders, this version of the heterotic string deserves increasing attention, even though it does not exhibit the protection that SUSY offers against the appearance of tachyons, quadratic divergences and a large cosmological constant [84].
The massless spectrum of this theory comprises three components: the gravitational part includes the graviton, the antisymmetric 2-form and the dilaton; the gauge bosons of arise in the gauge sector; and the charged matter states build the representations .
Applying similar compactification techniques such as orbifolds as in the supersymmetric case, there has been some effort to study the phenomenology of compactifications of this string theory in 4D, including models with a tachyon-free GUTs or SM massless spectrum [85, 86, 87, 88]. Although the progress does not yet compare to the supersymmetric case, some general features in SM-like models are known. In particular, the following properties of the massless spectrum are found: (i) at perturbative level, tachyons can be avoided; (ii) models with only one SM Higgs exist, but most of them exhibit a larger number of Higgses; (iii) there appear many fermion and scalar exotic states although there are models with a very small exotic spectrum; (iv) among the exotics, there are right-handed neutrinos; (v) leptoquark scalars are present in different amounts; and (vi) the number of fermions and bosons can coincide, yielding the possibility of an exponentially suppressed one-loop cosmological constant.
As an example, let us focus on the model 2 of [88], based on an Abelian orbifold compactification of the string (in the bosonic formulation). It includes the SM gauge group and additional and factors. In the fermionic sector, there are only three SM generations arising from twisted sectors and 119 right-handed neutrinos. In the scalar sector, besides 9 Higgs doublets and 9 scalar leptoquarks, there are 30 SM singlets, which may be considered flavons of a traditional flavor symmetry. The modular flavor features of this kind of models are not known.
8 Smooth compactifications
As already mentioned, one can obtain smooth compactifications of the heterotic theory in 10 dimensions. If the compactification is to preserve supersymmetry in 4 dimensions, the compact space has to be a CY manifold [5]. Models with the chiral spectrum of the MSSM have been found in this approach, see e.g. [35]. Notice that, a priori, it is not clear that every supergravity compactification of this type has a stringy origin [89], but is expected that a substantial fraction of the models in the literature correspond to string models. Machine learning techniques have been utilized to efficiently find models with the gauge symmetry and chiral spectrum of the SM [90]. It will be interesting to see if the absence of certain terms [91, 92] can be understood in terms of ordinary symmetries as is the case in the orbifold models discussed here, or if novel mechanisms are at play. In the latter case, this may provide us with new ways of overcoming C3. In passing, let us mention that a significant amount of smooth models can be obtained from orbifolds via blow-up (cf. e.g. [93]). In particular, giving VEVs to fields that are massless at the orbifold point often amounts to resolving the orbifold singularities. A detailed discussion of these interesting topics is, however, beyond the scope of this review.
9 Where do we stand?
The aim of heterotic model building is to reproduce and interpret particle physics in the heterotic string. This can be achieved by identifying appropriate compactifications. As we have discussed, various approaches have led to large sets of semi-realistic models that exhibit the matter spectrum of the standard model, its minimal supersymmetric version, as well as certain gauge extensions such as GUTs. Using various techniques, the effective symmetries of these constructions have been studied, which has yielded interesting implications for flavor physics, violation, proton stability, supersymmetry breaking, among other features.
However, a clear, let alone unique, picture has not yet emerged. The gauge and Yukawa couplings are, in principle, consistent with observation in a subset of the models. However, solid and precise predictions of the latter have remained largely elusive so far. This is hardly surprising. To see why, recall that we believe to know the Lagrange density of QCD in great detail but it remains a challenge to precisely compute basic quantities like the proton mass. In string phenomenology, the analogous analyses are even more challenging as the computation of many observables requires, among other things, a precise, quantitive understanding of moduli stabilization, which has not yet obtained. However, one can turn this around by saying that the explicit models provide us with a framework in which progress in these open questions can lead to testable predictions for the many parameters of the SM as well as BSM physics. We expect that this framework will also deliver a picture to address some of the pressing puzzles in cosmology.
- BBN
- big bang nucleosynthesis
- BSM
- beyond the standard model
- CFT
- conformal field theory
- CY
- Calabi–Yau
- EFT
- effective field theory
- FCNC
- flavor changing neutral current
- FFF
- Free Fermionic Formulation
- FI
- Fayet–Iliopoulos [2]
- GGSO
- Generalized GSO Projections
- GS
- Green–Schwarz [3]
- GGSO
- Generalized Gliozzi–Scherk–Olive [Gliozzi:1976qd] (GSO)
- GSO
- Gliozzi–Scherk–Olive [Gliozzi:1976qd]
- GUT
- Grand Unified Theory
- KK
- Kaluza–Klein
- LHC
- Large Hadron Collider
- MSSM
- minimal supersymmetric standard model
- PS
- Pati–Salam [1]
- QFT
- quantum field theory
- SB
- symmetry based
- SM
- standard model
- SUSY
- supersymmetry
- UV
- ultraviolet
- VEV
- vacuum expectation value
Acknowledgments
We would like to thank Carlo Angelantonj and Ignatios Antoniadis, and all the editors of the Handbook on Quantum Gravity, for inviting us to write this review. Big thanks go to our collaborators on this topic Wilfried Buchmüller, Mu-Chun Chen, Maximilian Fallbacher, Maximilian Fischer, Stefan Groot-Nibbelink, Koichi Hamaguchi, Rolf Kappl, Víctor Knapp-Pérez, Oleg Lebedev, Xiang-Gan Liu, Hans Peter Nilles, Yessenia Olguín-Trejo, Susha Parameswaran, Ricardo Pérez-Martínez, Felix Plöger, Stuart Raby, Graham Ross, Fabian Ruehle, Andreas Trautner, Patrick Vaudrevange, and Ivonne Zavala. The work of SRS is partially supported by UNAM-PAPIIT IN113223, CONACYT grant CB-2017-2018/A1-S-13051 and Marcos Moshinsky Foundation. The work of MR is supported by National Science Foundation grants PHY-1915005 and PHY-2210283, and parts of the work by MR were performed at the Aspen Center for Physics, which is supported by National Science Foundation grant PHY-160761. The authors were also supported by UC-MEXUS-CONACyT grant No. CN-20-38.
References
- [1] J. C. Pati and A. Salam, “Lepton Number as the Fourth Color,” Phys. Rev. D 10 (1974) 275–289. [Erratum: Phys.Rev.D 11, 703–703 (1975)].
- [2] P. Fayet and J. Iliopoulos, “Spontaneously Broken Supergauge Symmetries and Goldstone Spinors,” Phys. Lett. B 51 (1974) 461–464.
- [3] M. B. Green and J. H. Schwarz, “Anomaly Cancellation in Supersymmetric D=10 Gauge Theory and Superstring Theory,” Phys. Lett. B 149 (1984) 117–122.
- [4] D. I. Olive, “Relations between grand unified and monopole theories,”. Invited talk given at Study Conf. on Unification of Fundamental Interactions II, Erice, Italy, Oct 6-14, 1981.
- [5] P. Candelas, G. T. Horowitz, A. Strominger, and E. Witten, “Vacuum Configurations for Superstrings,” Nucl. Phys. B 258 (1985) 46–74.
- [6] L. J. Dixon, J. A. Harvey, C. Vafa, and E. Witten, “Strings on Orbifolds,” Nucl. Phys. B 261 (1985) 678–686.
- [7] L. J. Dixon, J. A. Harvey, C. Vafa, and E. Witten, “Strings on Orbifolds. 2.,” Nucl. Phys. B 274 (1986) 285–314.
- [8] M. Blaszczyk, S. Groot Nibbelink, F. Ruehle, M. Trapletti, and P. K. S. Vaudrevange, “Heterotic MSSM on a Resolved Orbifold,” JHEP 09 (2010) 065, arXiv:1007.0203 [hep-th].
- [9] L. E. Ibáñez and A. M. Uranga, String theory and particle physics: An introduction to string phenomenology. Cambridge University Press, 2, 2012.
- [10] P. Langacker, The Standard Model and Beyond. Taylor & Francis, 2017.
- [11] D. Baumann, Cosmology. Cambridge University Press, 7, 2022.
- [12] P. Langacker and G. Steigman, “Requiem for an FCHAMP? Fractionally CHArged, Massive Particle,” Phys. Rev. D 84 (2011) 065040, arXiv:1107.3131 [hep-ph].
- [13] J. Halverson and P. Langacker, “TASI Lectures on Remnants from the String Landscape,” PoS TASI2017 (2018) 019, arXiv:1801.03503 [hep-th].
- [14] S. P. Martin, “A Supersymmetry primer,” Adv. Ser. Direct. High Energy Phys. 18 (1998) 1–98, arXiv:hep-ph/9709356.
- [15] J. E. Kim and H. P. Nilles, “The mu Problem and the Strong CP Problem,” Phys. Lett. B 138 (1984) 150–154.
- [16] G. F. Giudice and A. Masiero, “A Natural Solution to the mu Problem in Supergravity Theories,” Phys. Lett. B 206 (1988) 480–484.
- [17] S. Dimopoulos, S. Raby, and F. Wilczek, “Supersymmetry and the Scale of Unification,” Phys. Rev. D 24 (1981) 1681–1683.
- [18] S. Raby, Supersymmetric Grand Unified Theories: From Quarks to Strings via SUSY GUTs, vol. 939. Springer, 2017.
- [19] J. D. Breit, B. A. Ovrut, and G. C. Segre, “E(6) Symmetry Breaking in the Superstring Theory,” Phys. Lett. B 158 (1985) 33.
- [20] C. Angelantonj and I. Florakis, “Introduction to String Theory,”. To appear as chapter of the Handbook of Quantum Gravity.
- [21] D. J. Gross, J. A. Harvey, E. J. Martinec, and R. Rohm, “The Heterotic String,” Phys. Rev. Lett. 54 (1985) 502–505.
- [22] R. Donagi and K. Wendland, “On orbifolds and free fermion constructions,” J. Geom. Phys. 59 (2009) 942–968, arXiv:0809.0330 [hep-th].
- [23] P. Athanasopoulos, A. E. Faraggi, S. Groot Nibbelink, and V. M. Mehta, “Heterotic free fermionic and symmetric toroidal orbifold models,” JHEP 04 (2016) 038, arXiv:1602.03082 [hep-th].
- [24] K. S. Narain, M. H. Sarmadi, and C. Vafa, “Asymmetric Orbifolds,” Nucl. Phys. B 288 (1987) 551.
- [25] L. E. Ibáñez, J. Mas, H.-P. Nilles, and F. Quevedo, “Heterotic Strings in Symmetric and Asymmetric Orbifold Backgrounds,” Nucl. Phys. B 301 (1988) 157–196.
- [26] D. Gepner, “New Conformal Field Theories Associated with Lie Algebras and their Partition Functions,” Nucl. Phys. B 290 (1987) 10–24.
- [27] D. Gepner, “Exactly Solvable String Compactifications on Manifolds of SU(N) Holonomy,” Phys. Lett. B 199 (1987) 380–388.
- [28] M. Fischer, M. Ratz, J. Torrado, and P. K. S. Vaudrevange, “Classification of symmetric toroidal orbifolds,” JHEP 01 (2013) 084, arXiv:1209.3906 [hep-th].
- [29] J. Giedt, “Spectra in standard - like Z(3) orbifold models,” Annals Phys. 297 (2002) 67–126, arXiv:hep-th/0108244.
- [30] F. Plöger, S. Ramos-Sánchez, M. Ratz, and P. K. S. Vaudrevange, “Mirage Torsion,” JHEP 04 (2007) 063, arXiv:hep-th/0702176.
- [31] C. Vafa, “Modular Invariance and Discrete Torsion on Orbifolds,” Nucl. Phys. B 273 (1986) 592–606.
- [32] D. Bailin and A. Love, “Orbifold compactifications of string theory,” Phys. Rept. 315 (1999) 285–408.
- [33] M. Fischer, S. Ramos-Sánchez, and P. K. S. Vaudrevange, “Heterotic non-Abelian orbifolds,” JHEP 07 (2013) 080, arXiv:1304.7742 [hep-th].
- [34] A. Hebecker and M. Trapletti, “Gauge unification in highly anisotropic string compactifications,” Nucl. Phys. B 713 (2005) 173–203, arXiv:hep-th/0411131.
- [35] V. Bouchard and R. Donagi, “An SU(5) heterotic standard model,” Phys. Lett. B 633 (2006) 783–791, arXiv:hep-th/0512149.
- [36] S. J. H. Konopka, “Non Abelian orbifold compactifications of the heterotic string,” JHEP 07 (2013) 023, arXiv:1210.5040 [hep-th].
- [37] H. P. Nilles and P. K. S. Vaudrevange, “Geography of Fields in Extra Dimensions: String Theory Lessons for Particle Physics,” Mod. Phys. Lett. A 30 no. 10, (2015) 1530008, arXiv:1403.1597 [hep-th].
- [38] Y. Olguín-Trejo, R. Pérez-Martínez, and S. Ramos-Sánchez, “Charting the flavor landscape of MSSM-like Abelian heterotic orbifolds,” Phys. Rev. D 98 no. 10, (2018) 106020, arXiv:1808.06622 [hep-th].
- [39] E. Parr and P. K. S. Vaudrevange, “Contrast data mining for the MSSM from strings,” Nucl. Phys. B 952 (2020) 114922, arXiv:1910.13473 [hep-th].
- [40] E. Witten, “Strong coupling expansion of Calabi-Yau compactification,” Nucl. Phys. B 471 (1996) 135–158, arXiv:hep-th/9602070.
- [41] K. R. Dienes, “String theory and the path to unification: A Review of recent developments,” Phys. Rept. 287 (1997) 447–525, arXiv:hep-th/9602045.
- [42] M. Quiros, “New ideas in symmetry breaking,” in Theoretical Advanced Study Institute in Elementary Particle Physics (TASI 2002): Particle Physics and Cosmology: The Quest for Physics Beyond the Standard Model(s), pp. 549–601. 2, 2003. arXiv:hep-ph/0302189.
- [43] H. P. Nilles, S. Ramos-Sánchez, P. K. S. Vaudrevange, and A. Wingerter, “The Orbifolder: A Tool to study the Low Energy Effective Theory of Heterotic Orbifolds,” Comput. Phys. Commun. 183 (2012) 1363–1380, arXiv:1110.5229 [hep-th].
- [44] L. McAllister and F. Quevedo, “Moduli Stabilization in String Theory,” arXiv:2310.20559 [hep-th]. To appear as chapter of the Handbook of Quantum Gravity.
- [45] L. J. Dixon, V. Kaplunovsky, and J. Louis, “On Effective Field Theories Describing (2,2) Vacua of the Heterotic String,” Nucl. Phys. B 329 (1990) 27–82.
- [46] M.-C. Chen, M. Fallbacher, K. T. Mahanthappa, M. Ratz, and A. Trautner, “CP Violation from Finite Groups,” Nucl. Phys. B 883 (2014) 267–305, arXiv:1402.0507 [hep-ph].
- [47] A. Baur, H. P. Nilles, A. Trautner, and P. K. S. Vaudrevange, “Unification of Flavor, CP, and Modular Symmetries,” Phys. Lett. B 795 (2019) 7–14, arXiv:1901.03251 [hep-th].
- [48] A. Baur, H. P. Nilles, A. Trautner, and P. K. S. Vaudrevange, “A String Theory of Flavor and ,” Nucl. Phys. B 947 (2019) 114737, arXiv:1908.00805 [hep-th].
- [49] K. S. Narain, “New Heterotic String Theories in Uncompactified Dimensions 10,” Phys. Lett. B 169 (1986) 41–46.
- [50] S. Groot Nibbelink and P. K. S. Vaudrevange, “T-duality orbifolds of heterotic Narain compactifications,” JHEP 04 (2017) 030, arXiv:1703.05323 [hep-th].
- [51] G. Senjanovic and R. N. Mohapatra, “Exact Left-Right Symmetry and Spontaneous Violation of Parity,” Phys. Rev. D 12 (1975) 1502.
- [52] S. M. Barr, “A New Symmetry Breaking Pattern for SO(10) and Proton Decay,” Phys. Lett. B 112 (1982) 219–222.
- [53] A. Font, L. E. Ibanez, H. P. Nilles, and F. Quevedo, “Degenerate Orbifolds,” Nucl. Phys. B 307 (1988) 109–129. [Erratum: Nucl.Phys.B 310, 764–764 (1988)].
- [54] F. Buccella, J. P. Derendinger, S. Ferrara, and C. A. Savoy, “Patterns of Symmetry Breaking in Supersymmetric Gauge Theories,” Phys. Lett. B115 (1982) 375.
- [55] G. Cleaver, M. Cvetic, J. R. Espinosa, L. L. Everett, and P. Langacker, “Intermediate scales, mu parameter, and fermion masses from string models,” Phys. Rev. D 57 (1998) 2701–2715, arXiv:hep-ph/9705391.
- [56] M. Cvetic, L. L. Everett, and J. Wang, “Units and numerical values of the effective couplings in perturbative heterotic string vacua,” Phys. Rev. D 59 (1999) 107901, arXiv:hep-ph/9808321.
- [57] R. Kappl, M. Ratz, and C. Staudt, “The Hilbert basis method for D-flat directions and the superpotential,” JHEP 10 (2011) 027, arXiv:1108.2154 [hep-th].
- [58] P. Binetruy and P. Ramond, “Yukawa textures and anomalies,” Phys. Lett. B 350 (1995) 49–57, arXiv:hep-ph/9412385.
- [59] J. P. Derendinger, L. E. Ibáñez, and H. P. Nilles, “On the Low-Energy d = 4, N=1 Supergravity Theory Extracted from the d = 10, N=1 Superstring,” Phys. Lett. B 155 (1985) 65–70.
- [60] K. A. Intriligator, N. Seiberg, and D. Shih, “Dynamical SUSY breaking in meta-stable vacua,” JHEP 04 (2006) 021, arXiv:hep-th/0602239.
- [61] B. Petersen, M. Ratz, and R. Schieren, “Patterns of remnant discrete symmetries,” JHEP 08 (2009) 111, arXiv:0907.4049 [hep-ph].
- [62] H. Ishimori, T. Kobayashi, H. Ohki, Y. Shimizu, H. Okada, and M. Tanimoto, “Non-Abelian Discrete Symmetries in Particle Physics,” Prog. Theor. Phys. Suppl. 183 (2010) 1–163, arXiv:1003.3552 [hep-th].
- [63] S. Ramos-Sánchez and P. K. S. Vaudrevange, “Note on the space group selection rule for closed strings on orbifolds,” JHEP 01 (2019) 055, arXiv:1811.00580 [hep-th].
- [64] F. Beye, T. Kobayashi, and S. Kuwakino, “Gauge Origin of Discrete Flavor Symmetries in Heterotic Orbifolds,” Phys. Lett. B 736 (2014) 433–437, arXiv:1406.4660 [hep-th].
- [65] G. Lopes Cardoso, D. Lüst, and T. Mohaupt, “Moduli spaces and target space duality symmetries in (0,2) Z(N) orbifold theories with continuous Wilson lines,” Nucl. Phys. B 432 (1994) 68–108, arXiv:hep-th/9405002.
- [66] S. Ferrara, D. Lüst, A. D. Shapere, and S. Theisen, “Modular Invariance in Supersymmetric Field Theories,” Phys. Lett. B 225 (1989) 363.
- [67] L. J. Dixon, D. Friedan, E. J. Martinec, and S. H. Shenker, “The Conformal Field Theory of Orbifolds,” Nucl. Phys. B 282 (1987) 13–73.
- [68] J. Lauer, J. Mas, and H. P. Nilles, “Duality and the Role of Nonperturbative Effects on the World Sheet,” Phys. Lett. B 226 (1989) 251–256.
- [69] J. Lauer, J. Mas, and H. P. Nilles, “Twisted sector representations of discrete background symmetries for two-dimensional orbifolds,” Nucl. Phys. B 351 (1991) 353–424.
- [70] L. E. Ibáñez and D. Lüst, “Duality anomaly cancellation, minimal string unification and the effective low-energy Lagrangian of 4-D strings,” Nucl. Phys. B 382 (1992) 305–361, arXiv:hep-th/9202046.
- [71] H. P. Nilles, S. Ramos–Sánchez, and P. K. S. Vaudrevange, “Eclectic flavor scheme from ten-dimensional string theory - II detailed technical analysis,” Nucl. Phys. B 966 (2021) 115367, arXiv:2010.13798 [hep-th].
- [72] H. P. Nilles, M. Ratz, A. Trautner, and P. K. S. Vaudrevange, “ violation from string theory,” Phys. Lett. B 786 (2018) 283–287, arXiv:1808.07060 [hep-th].
- [73] S. Biermann, A. Mütter, E. Parr, M. Ratz, and P. K. S. Vaudrevange, “Discrete remnants of orbifolding,” Phys. Rev. D 100 no. 6, (2019) 066030, arXiv:1906.10276 [hep-ph].
- [74] R. Kappl, H. P. Nilles, S. Ramos-Sánchez, M. Ratz, K. Schmidt-Hoberg, and P. K. S. Vaudrevange, “Large hierarchies from approximate R symmetries,” Phys. Rev. Lett. 102 (2009) 121602, arXiv:0812.2120 [hep-th].
- [75] K.-S. Choi, H. P. Nilles, S. Ramos-Sánchez, and P. K. S. Vaudrevange, “Accions,” Phys. Lett. B 675 (2009) 381–386, arXiv:0902.3070 [hep-th].
- [76] M. Dine, A. E. Nelson, Y. Nir, and Y. Shirman, “New tools for low-energy dynamical supersymmetry breaking,” Phys. Rev. D 53 (1996) 2658–2669, arXiv:hep-ph/9507378.
- [77] S. Ramos-Sánchez, M. Ratz, Y. Shirman, S. Shukla, and M. Waterbury, “Generation flow in field theory and strings,” JHEP 10 (2021) 144, arXiv:2109.01681 [hep-th].
- [78] N. Sakai and T. Yanagida, “Proton Decay in a Class of Supersymmetric Grand Unified Models,” Nucl. Phys. B 197 (1982) 533.
- [79] G. Altarelli and F. Feruglio, “SU(5) grand unification in extra dimensions and proton decay,” Phys. Lett. B 511 (2001) 257–264, arXiv:hep-ph/0102301.
- [80] R. Kappl, B. Petersen, S. Raby, M. Ratz, R. Schieren, and P. K. S. Vaudrevange, “String-Derived MSSM Vacua with Residual R Symmetries,” Nucl. Phys. B 847 (2011) 325–349, arXiv:1012.4574 [hep-th].
- [81] I. Antoniadis, E. Gava, K. S. Narain, and T. R. Taylor, “Effective mu term in superstring theory,” Nucl. Phys. B 432 (1994) 187–204, arXiv:hep-th/9405024.
- [82] L. Alvarez-Gaumé, P. H. Ginsparg, G. W. Moore, and C. Vafa, “An O(16) x O(16) Heterotic String,” Phys. Lett. B 171 (1986) 155–162.
- [83] L. J. Dixon and J. A. Harvey, “String Theories in Ten-Dimensions Without Space-Time Supersymmetry,” Nucl. Phys. B 274 (1986) 93–105.
- [84] S. Groot Nibbelink, O. Loukas, A. Mütter, E. Parr, and P. K. S. Vaudrevange, “Tension Between a Vanishing Cosmological Constant and Non-Supersymmetric Heterotic Orbifolds,” Fortsch. Phys. 68 no. 7, (2020) 2000044, arXiv:1710.09237 [hep-th].
- [85] M. Blaszczyk, S. Groot Nibbelink, O. Loukas, and S. Ramos-Sánchez, “Non-supersymmetric heterotic model building,” JHEP 10 (2014) 119, arXiv:1407.6362 [hep-th].
- [86] S. Abel, K. R. Dienes, and E. Mavroudi, “Towards a nonsupersymmetric string phenomenology,” Phys. Rev. D 91 no. 12, (2015) 126014, arXiv:1502.03087 [hep-th].
- [87] S. Abel, K. R. Dienes, and E. Mavroudi, “GUT precursors and entwined SUSY: The phenomenology of stable nonsupersymmetric strings,” Phys. Rev. D 97 no. 12, (2018) 126017, arXiv:1712.06894 [hep-ph].
- [88] R. Pérez-Martínez, S. Ramos-Sánchez, and P. K. S. Vaudrevange, “Landscape of promising nonsupersymmetric string models,” Phys. Rev. D 104 no. 4, (2021) 046026, arXiv:2105.03460 [hep-th].
- [89] S. Groot Nibbelink, O. Loukas, F. Ruehle, and P. K. S. Vaudrevange, “Infinite number of MSSMs from heterotic line bundles?,” Phys. Rev. D 92 no. 4, (2015) 046002, arXiv:1506.00879 [hep-th].
- [90] F. Ruehle, “Data science applications to string theory,” Phys. Rept. 839 (2020) 1–117.
- [91] E. Silverstein and E. Witten, “Criteria for conformal invariance of (0,2) models,” Nucl. Phys. B 444 (1995) 161–190, arXiv:hep-th/9503212.
- [92] L. B. Anderson, J. Gray, M. Larfors, and M. Magill, “Vanishing Yukawa Couplings and the Geometry of String Theory Models,” arXiv:2201.10357 [hep-th].
- [93] S. Groot Nibbelink, J. Held, F. Ruehle, M. Trapletti, and P. K. S. Vaudrevange, “Heterotic Z(6-II) MSSM Orbifolds in Blowup,” JHEP 03 (2009) 005, arXiv:0901.3059 [hep-th].
- BBN
- big bang nucleosynthesis
- BSM
- beyond the standard model
- CFT
- conformal field theory
- CY
- Calabi–Yau
- EFT
- effective field theory
- FCNC
- flavor changing neutral current
- FFF
- Free Fermionic Formulation
- FI
- Fayet–Iliopoulos [2]
- GGSO
- Generalized GSO Projections
- GS
- Green–Schwarz [3]
- GGSO
- Generalized GSO
- GSO
- Gliozzi–Scherk–Olive [Gliozzi:1976qd]
- GUT
- Grand Unified Theory
- KK
- Kaluza–Klein
- LHC
- Large Hadron Collider
- MSSM
- minimal supersymmetric standard model
- PS
- Pati–Salam [1]
- QFT
- quantum field theory
- SB
- symmetry based
- SM
- standard model
- SUSY
- supersymmetry
- UV
- ultraviolet
- VEV
- vacuum expectation value