Δ\frozen@everymath\macc@group\macc@set@skewchar\macc@nested@a111|
Projective Superspaces I:
Interacting Linear Sigma Models
We describe the projective superspace approach to supersymmetric models with off-shell supersymmetry in two dimensions. In addition to the usual superspace coordinates, projective superspace has extra bosonic variables – one doublet for each in the R-symmetry which are interpreted as homogeneous coordinates on . The superfields are analytic in the coordinates and this analyticity plays an important role in our description. For instance, it leads to stringent constraints on the interactions one can write down for a given superfield content of the model. As an example, we describe in projective superspace Witten’s ADHM sigma model – a linear sigma model with non-derivative interactions whose target is with a Yang-Mills instanton solution. The hyperkähler nature of target space and the twistor description of instantons by Ward, and Atiyah, Hitchin, Drinfeld and Manin are natural outputs of our construction.
1 Introduction
Two dimensional quantum field theories with chiral supersymmetry have appeared in a variety of physical and mathematical contexts. The most familiar example is the construction of heterotic string models which have supersymmetry on the worldsheet [GHMR85]. Conformal theories with supersymmetry were explored [CHSW85, HW85] in the context of compactifications of the type where is a compact Calabi-Yau threefold. Landau-Ginzburg models were also found to furnish a large class of heterotic sigma models [DK94]. worldsheet conformal theories are also interesting: they describe compactifications to six dimensions [BD88, EOTY89, Sei88] and are useful in worldsheet descriptions of five-brane instantons [CHS91, CHS91a].
Since the brane revolution, many two dimensional spacetime models with chiral supersymmetry have been constructed – these appear as low-energy effective theories on two dimensional intersections of D-branes or on D1-branes probing manifolds with special holonomy. Depending on the brane setup, the models on the intersection may have , , or even supersymmetry [GHM97, BDL96]. Typically, D-branes have gauge fields as part of their low-energy dynamics and the chiral supersymmetric theory is a gauged linear sigma model.
For example, a D1-brane probing a manifold has supersymmetry on its worldvolume whereas it has supersymmetry when probing a Calabi-Yau fourfold. The intersection of two stacks of D5-branes on a two-dimensional plane has supersymmetry on the common intersection [IKS06]; including a probe D1-brane on the common intersection gives susy on the intersection [GMMS05, Ton14, Nek16, NP17]. Another system of D-branes which has susy is the system which is a D1-brane probe of a gauge theory instanton on realized by the system [Dou98], or instantons on an ALE space realized by taking the four transverse directions of the D9-brane relative to the D5-brane [DM96]. Other brane realizations include the worldvolume theory on M5-branes wrapped on a coassociative submanifold of a -manifold which has supersymmetry [GGP16] and M5-branes wrapped on a four dimensional submanifold of a Calabi-Yau threefold which has supersymmetry [MSW97, GGP14, PSY16].
Superspace has proven to be powerful in understanding supersymmetric theories primarily because it realizes the supersymmetry algebra off-shell. The advantage of an off-shell realization is that, as long as the constraints on superfields do not themselves introduce interactions, we have a clean separation of kinematics and dynamics and the sum of two supersymmetric actions is automatically supersymmetric. This has been useful in uncovering the geometric structures hidden in supersymmetric theories and also understanding dualities between very different-looking models [LR83, HKLR87]. However, the presence of so-called -terms can mix dynamics with kinematics, and then supersymmetry restricts the structure of the action even in superspace; we shall see that this plays a crucial role in our construction of interacting models.
Superspace descriptions of , , , and theories exist [Sak85, Sie84, FM73, BW76, Zum75, DS86, Ade+76, BMG86] and are well-understood. For theories with a higher amount of supersymmetry, for instance in two dimensions (more generally, theories with eight supercharges in other dimensions), it is well known that ordinary superspace is not sufficient to describe off-shell charged hypermultiplets since the superspace constraints for the hypermultiplet put it on-shell (see [GGRS83, Section 4.6]).
There are at least two approaches that address these issues, harmonic superspace [GIKO+84, GIOS84] and the closely related isotwistor superspace [Ros83, RS86, Ros85], and projective superspace [GHR84, KLR84]. All approaches introduce a new set of bosonic coordinates which are coordinates on an . In the harmonic approach the are viewed as harmonic coordinates on where is the R-symmetry group or a subgroup thereof, and one considers superfields which are harmonic functions on . In the projective approach, the is viewed as and the are homogeneous coordinates on the and the superfields are analytic functions on . These two approaches are in fact related [Kuz99, JS09].
Projective superspace has been successful in describing many supersymmetric models with eight supercharges [KLR84, KLR87, IR96, LR88, LR90, LR10, AKL07, GL85, GK99]. In projective superspace, one can write down new kinds of superfields and superspace constraints which depend on the coordinates . More precisely, they are fibred over the coset space . Superfields over projective superspace typically contain an infinite number of ordinary superfields (the coefficients in a Taylor expansion in ) and these turn out to be crucial in realizing the off-shell version of the hypermultiplet. Dynamically, most of these superfields turn out to be auxiliary and thus do not change the on-shell content of the hypermultiplet.
projective superspace has been introduced in [HL17, HL17a] and has been used to give off-shell formulations of nonlinear sigma models involving hypermultiplets. In this paper, we describe linear sigma models with manifest off-shell supersymmetry in projective superspace.
The R-symmetry of the supersymmetry algebra is and thus one has two projective superspaces with the s corresponding to the two subgroups. The hypermultiplets are also of two kinds, transforming as a doublet under either or . We call them standard hypermultiplets and, following [Wit95], twisted hypermultiplets respectively. We describe these in detail in Section 3. We shall see that a hyper can be realized either as a linear polynomial in the homogeneous coordinates (the superfield) or as a power series in a local coordinate on the (the ‘polar’ superfield). superfields are treated in some detail in [HL17, HL17a]. supersymmetry allows independent fermionic multiplets with chirality opposite to that of the supercharges. These are the fermi multiplets; we realize them in projective superspace in Section 4.
In models, we have interactions of the nonlinear sigma model type or the non-derivative type. Non-derivative interactions between chiral multiplets, gauge multiplets and fermi multiplets are described by modifying their superspace constraints with the so-called -terms, or by including superpotential-like -terms in the Lagrangian (see Appendix A.2 of this paper). In Section 5, we describe the -term type non-derivative interactions for models containing standard hypers, twisted hypers and fermis (it turns out that -terms are not possible). In a companion paper [PR], we describe gauge multiplets and their interactions with hypers and fermis in projective superspace.
In Section 6, we describe in projective superspace a prominent supersymmetric model: a linear sigma model which flows down to a sigma model with target being an instanton solution in four dimensions. The couplings of the linear sigma model and the constraints they satisfy as a consequence of supersymmetry encode the data that enters ADHM construction of instantons [AHDM78]. This was demonstrated in superspace by Witten [Wit95], and it was given a D-brane interpretation by Douglas [Dou98]. A manifest construction was given in harmonic superspace in [GS95, GS96] (see [GR95] for some partial results in ordinary superspace). In our construction in projective superspace, the hyperkähler nature of the target space is manifest, and the monads which describe holomorphic bundles on twistor space [AHDM78] appear explicitly. We also extend this construction to self-dual solutions on with , i.e., of dimension greater than 4 [CGK85, Wit95].
The appendix includes a quick review of and superspaces (Appendix A), a realization of the hypermultiplet in projective superspace and its reduction to projective superspace (Appendix B), and finally a detailed derivation of the ordinary space component actions for the general supersymmetric interacting linear sigma model (Appendix C).
2 projective superspace
2.1 Introduction
The supersymmetric algebra consists of four real supercharges , , of right-handed chirality. It is useful to write these real supercharges in terms of a matrix that satisfies the reality conditions
(2.1) |
where . Here, and are -doublet indices. The R-symmetry group is then . We will be interested in the representations of the supersymmetry algebra which are charged under just one of the s and hence it is useful to consider the double cover . The and indices are lowered using the invariant tensors and which satisfy , and .
The supersymmetry algebra is
(2.2) |
superspace is described by the supercoordinates where . The corresponding supercovariant derivatives are and with the algebra
(2.3) |
The derivatives also satisfy the same reality condition as for the supersymmetry generators (2.1). We loosely refer to (2.3) as the supersymmetry algebra though it differs from (2.2) by a sign. The supersymmetry generators and the derivatives mutually anticommute: .
In this paper, we work exclusively with the derivatives rather than the supersymmetry generators . Supersymmetry transformations of some component of a superfield can be expressed in terms of because of the following fact which can be easily verified by using the explicit superspace expressions for and :
(2.4) |
where are constant Grassmann parameters, and stands for the operation of setting the Grassmann coordinates to zero in the expression . The that appear in the superspace measure can also replaced by the corresponding up to total derivatives333This is standard procedure, see e.g. [GGRS83]..
It is convenient to define
(2.5) |
These derivatives span two (anti)commuting subalgebras:
(2.6) |
2.2 Algebras, superfields and actions
Consider two sets of commuting coordinates and which are doublets under the R-symmetry subgroups and respectively. These are most usefully interpreted in our context as homogeneous coordinates on (we label the second as to indicate its relation to ). The superspace with the coordinates is which we refer to as projective superspace. The subspaces and are important for us.
We also introduce conjugate doublets and which satisfy
(2.7) |
A shift symmetry
Note that there is more than one solution to the equation . If is one solution, then so is for any . Thus there is a shift symmetry on the :
(2.8) |
There is a similar shift symmetry for the conjugate doublet .
Derivatives on projective superspace
Consider the derivatives
(2.9) |
where and are any solutions to the equations (2.7). The algebra of the derivatives (2.9) is obtained from (2.3):
(2.10) |
Note that the shift symmetry (2.8) shifts the derivatives by but it leaves the algebra (2.2) unchanged. We shall see below that the action is also invariant under the shift symmetry up to total derivative terms.
Projective superfields
An -projective superfield is a function of the superspace coordinates and the coordinates which satisfy the following:
-
(1)
is holomorphic in a domain in ,
-
(2)
satisfies the projective constraints ,
-
(3)
may be in non-trivial representations of the R-symmetry group and the Lorentz group .
An -projective superfield is analogously a function of the superspace coordinates and the coordinate and is annihilated by . We discuss the different types of projective superfields in Section 2.4.
The -projective constraints can be encoded more economically in terms of the fully contracted derivative (2.11) :
(2.12) |
Since depends only on and not on , implies . The advantage of (2.12) is that it takes the same form for -projective superfields as well, since we can now recover using . We frequently use the derivative in the paper.
Actions
The constraints on a projective superfield imply that depends on only half of the Grassmann coordinates. The appropriate superspace measure which ensures invariance of an action composed of projective superfields is then quadratic in the derivatives , i.e., . The supersymmetric action is then given by
(2.13) |
where
-
1.
sets all the Grassmann coordinates to zero (we frequently omit the from our expressions).
-
2.
is the superspace Lagrangian which satisfies . It must carry the Lorentz representation (left-moving part of a vector) in order to compensate the in the projective superspace measure.
-
3.
The contour is chosen to avoid possible singularities in .
The action is invariant (up to total spacetime derivatives) under the shift symmetry (2.8) since the Lagrangian satisfies . Since the superspace measure is invariant under and , the action (2.13) is manifestly invariant under and if the superspace Lagrangian is invariant.
Non-derivative interactions
Suppose a projective superfield is in the spin representation of the Lorentz group . The requirement that can be relaxed to have a non-zero right hand side:
(2.14) |
where is a function of other superfields in the model and is in the spin representation of . This allows us to introduce interactions (the so-called -terms) as we will see later in Section 5:
The modified constraints (2.14) are consistent with the algebra only if the function satisfies
(2.15) |
To ensure invariance of the action, we require that the superspace Lagrangian satisfies even if is not zero. This further constrains the .
Thus, any supersymmetric model must satisfy the following constraints:
-
1.
The algebra must be satisfied on every superfield in the model,
-
2.
The superspace Lagrangian must satisfy to ensure supersymmetry of the action.
These criteria place stringent constraints on the superfield content and the interactions in a model.
2.3 Projective superspace in inhomogeneous coordinates
A primer on
The projective space is constructed as the quotient space , where is the following equivalence relation on the coordinates of : , . We describe in terms of two charts and :
(2.16) |
The map which acts on the homogeneous coordinates as
(2.17) |
interchanges the two charts. Using the equivalence , , we can scale out the non-zero coordinate in each of the charts and obtain a description in terms of inhomogeneous coordinates:
(2.18) |
with and . On the intersection , the local coordinates and are related by the map (2.17)
(2.19) |
We can thus express all our results exclusively in terms of one of the inhomogeneous coordinates, say , by appending the point to the chart . We frequently adopt this usage to avoid cluttering of notation.
The derivatives ,
We next express the derivatives and in terms of the local coordinates and in the charts and respectively. In the chart , we have and . Thus, we can choose which indeed satisfies . Using the scale invariance , we can set as discussed above (2.18). The derivatives and are then given by
(2.20) |
where we have used the expressions (2.5) for . A similar description can be obtained in the chart in which . Writing , choosing and setting by scale invariance, we have
(2.21) |
Observe that, in the intersection , the derivatives defined in are related to the defined in as
(2.22) |
which is the gluing rule for a global section of the line bundle (we have used that on the overlap).
Similarly, the derivatives in and in are related on the overlap as
(2.23) |
where we have used (2.22) in going to the last expression. The transformation (2.23) can be viewed as the usual transformation of a section of plus a shift term proportional to generated by the shift symmetry (2.8). This allows us to define globally on , not as a section of but as a section of the affine bundle modelled on .
Note: There is an alternate way of writing the algebra using the derivatives and :
(2.24) |
Thus, one may use the derivatives and instead of and in describing projective superspace. Observe that coincides with in and coincides with in . Further, also satisfies the rule (2.23). However, this is expected since the derivative of a section of transforms as a section of the affine bundle modelled on .
A action which is supersymmetric
-projective superspace
For completeness, we explicitly describe some analogous aspects of -projective superspace. We have the inhomogeneous coordinate for the corresponding to the doublet . We then choose and using the scale invariance , . The -projective derivatives and are then
(2.27) |
A supersymmetric action in superspace for -projective superfields is given by
(2.28) |
The actions we consider in this paper will only have a single contour integral over either or .
The fully contracted derivative (2.11) in terms of and is
(2.29) |
2.4 Analytic structure of projective superfields
Recall that -projective superfields are holomorphic in a connected open subset of and that they are annihilated by the derivatives . We now describe the different types of projective superfields which differ in their analytic structure on the . -projective superfields are defined analogously.
superfields
The superfield is a homogeneous polynomial in the of degree :
(2.30) |
The components are ordinary superfields, i.e., functions on . Note that is a global section of the line bundle . We thus call such superfields superfields. In the chart where we can write as
(2.31) |
which becomes a polynomial in the inhomogeneous coordinate when we set .
Meromorphic superfields
The superfields discussed above are global holomorphic sections of . We can consider more general superfields which are only local sections of and cannot be extended to all of . A familiar class of examples are the meromorphic sections of which are rational functions of :
(2.32) |
where and are homogeneous polynomials of degree and respectively. The domain of definition of on is restricted to the open set where . The degree of homogeneity of is then and thus is a local section of defined on . In terms of the inhomogeneous coordinate , we have
(2.33) |
where the are appropriate combinations of and similarly, are combinations of the .
Local superfields
Consider superfields which are formal power series in or or both. These appear as series expansions of local holomorphic sections in the neighbourhoods of , or in the annulus . Consider a power series in :
(2.34) |
Such superfields shall be termed arctic since they are well-defined at the north pole of (and possibly in a neighbourhood of as well). Similarly, a superfield which is a power series in is designated antarctic.
Finally, a superfield which is defined in the annulus and is real under the extended complex conjugation given below in Section 2.6 is called equatorial.
2.5 R-symmetry in projective superspace
We consider the R-symmetry transformation of the various objects in projective superspace for the subgroup in this subsection [KLR84] (the discussion for proceeds analogously). The homogeneous coordinates on transforms as a doublet under :
(2.35) |
Accordingly, the inhomogeneous coordinate transforms fractional-linearly:
(2.36) |
Also, a doublet with a lower index transforms as
(2.37) |
Factor of automorphy
We define a factor of automorphy for the action of on as follows. Let and . Then we have
(2.38) |
It is easy to check that satisfies . Suppose we have an object that depends holomorphically on . The transformation of by a -transformation is denoted by . An object is said to have -weight if it satisfies
(2.39) |
That is, is a local section of the line bundle . Note that weight objects are simply local functions on .
Next, we describe the -symmetry of superfields and arctic superfields.
superfields
Consider an superfield given by . Since all -doublet indices are contracted in , it is invariant under . That is,
(2.40) |
where on the right hand side is a new superfield obtained by transforming the components , and evaluated at the transformed coordinates . In terms of the inhomogeneous coordinate , we have
(2.41) |
where are appropriate combinations of the . Similarly,
(2.42) |
This leads to
(2.43) |
We define the transformation of a superfield by an element as
(2.44) |
where the right hand side must be expanded about so that it is a function of rather than . Thus, an superfield has weight (note that this is also the degree of the line bundle ). Meromorphic sections of also transform similarly under R-symmetry.
An example
We are primarily interested in describing hypermultiplets which correspond to . In this case the components of transform as an -doublet. We check that satisfies (2.43) for :
(2.45) |
where we have used the transformation of a doublet with a lower index as described in eq. (2.37). It can be easily checked that the conjugate (cf. (2.64)) also transforms as an multiplet.
Arctic superfields
Arctic superfields are typically defined only in a neighbourhood of and not globally on . As a result, we may only consider infinitesimal R-symmetry transformations of arctic superfields since they retain in a neighbourhood of . These we obtain by setting and , with and infinitesimal, in the formula for the -transformation in (2.35). The determinant condition then gives to first order in the infinitesimals, i.e., is real. The infinitesimal -transformation of is then (cf. [LR10])
(2.46) |
The -transformation of an arctic superfield of weight is then given by the infinitesimal version of :
(2.47) |
It is important to note that arctic superfields can be assigned any integral weight a priori since arctics go to arctics under infinitesimal transformations for any in eq. (2.47)444Explicitly, we have Clearly, the right hand side is also an arctic superfield.. Further, it is easy to check that also transforms as an object of weight but is no longer an antarctic superfield.
The components transform under (2.47) as
(2.48) |
Let us look at which will be required in our study of hypermultiplets. We shall show below that, with our choice of action for the arctic superfield, the components will turn out to be auxiliary and will be set to zero by their equations of motion. The arctic superfield then truncates to an superfield after substituting . It is then clear that the components and decouple from the components in (2.48) and and transform as
(2.49) |
These are the transformation rules for an -doublet and this is the standard transformation of a hypermultiplet under R-symmetry.
The derivatives ,
Since , the same manipulations we did for superfields works here and it follows from (2.43) that has -weight . Let us next discuss the -weight of . Recall from the discussion above equation (2.20) that our chosen solution for the equation is
(2.50) |
Under -transformations, since transforms as , and transform as
(2.51) |
From this, it is clear that has -weight . This is consistent with the algebra since the right hand side is independent of and hence, has weight .
However, the transformation (2.51) of does not look like that of an -doublet. The latter looks like
(2.52) |
How do we reconcile (2.51) and (2.52)? Recall that we had a shift symmetry (2.8) in the space of that satisfy . We could add a shift in one of the transformations, say (2.51) and see if that can be matched with (2.52) for a particular value of the shift parameter. Indeed, writing
(2.53) |
we get a solution for
(2.54) |
In analogy with (2.44), we define the transformations of the and expressed in inhomogeneous coordinates as
(2.55) |
The projective superspace measure
Recall that the projective superspace action (2.13) is
(2.56) |
As discussed after (2.13), the action is manifestly and invariant provided the superspace Lagrangian is invariant. In terms of -weight, it has weight since the measure has two factors of and has two factors of . Let us elaborate in terms of inhomogeneous coordinates. The action takes the form
(2.57) |
The measure transforms with -weight under an -transformation :
(2.58) |
The superderivatives effectively transform with -weight (cf. the first term in the transformation of in (2.55); the second term in (2.55) is proportional to which annihilates ). As a result, the combination has weight , i.e., the superspace measure is invariant (up to total derivatives). Since integrating a weight object with the invariant measure yields an -invariant answer, the action is R-symmetric if the superspace Lagrangian has weight .
2.6 Extended complex conjugation
Recall the map (2.17) which takes . The antipodal map that takes a point on to its antipode is the composition of the map and complex conjugation:
(2.59) |
The antipodal map can be used to define a new real structure [LR88] on the (sheaf of) sections of a line bundle as the action of the antipodal map on a section followed by ordinary complex conjugation of the resulting section.
For instance, the antipodal map acts on an arctic superfield (which is a local section of some line bundle on ) as
(2.60) |
Ordinary complex conjugation of the resulting local section is
(2.61) |
Thus, the extended complex conjugate of an arctic superfield is
(2.62) |
Let us compute the extended complex conjugate of an superfield . Since is globally defined on , and the antipodal map contains the map, we can use the -transformation rule (2.44) for superfields to obtain the extended complex conjugate:
(2.63) |
The difference between the above and (2.62) is that there is an additional factor of in the antipodal map step. This factor makes the new section also a global section of . Thus, the extended complex conjugate of an superfield is
(2.64) |
A reality condition
As is obvious from (2.64), the extended conjugate of an superfield is also an superfield. Notice that applying the extended complex conjugate twice on gives
(2.65) |
Thus, we can impose a reality condition on an superfield only when is even:
(2.66) |
Extended complex conjugates of and
Next, consider the derivatives and . Since they are globally defined (see the equations (2.22), (2.23) and the discussion around them), we use the global -transformation rules in (2.55) to get the conjugates. The factor of automorphy for the -map is . The complex conjugate of is then
(2.67) |
The complex conjugate of is obtained as follows. First, we apply the antipodal map:
(2.68) |
where we have used the fact that since the are independent of , the expressions for are the same as in (2.20), i.e., . Next, doing ordinary complex conjugation, we get
(2.69) |
Thus, we have
(2.70) |
We may need to consider a slightly different version of the complex conjugates of the derivatives when they act on arctic superfields for the following reason. Under extended complex conjugation, an arctic superfield goes to an antarctic superfield (see (2.62)). We would like this to be true for the derivative of an arctic as well. However, applying (2.70) on gives which is not antarctic due to a term proportional to in . On the other hand, treating as a new arctic superfield with components
(2.71) |
we can apply the conjugation rule (2.62) to the above and obtain
(2.72) |
as the conjugate antarctic superfield corresponding to . Clearly, (2.72) can be written as
(2.73) |
which suggests that we modify the conjugate of the derivative when acting on arctic superfields to
(2.74) |
Similarly, we have
(2.75) |
Applying (2.62) to the above, we get
(2.76) |
Note the identity
(2.77) |
This allows us to write (2.76) as
(2.78) |
which suggests the modification
(2.79) |
Thus, on arctic superfields, we can postulate the following modified extended complex conjugates of the derivatives:
(2.80) |
The notation for the above notion of the extended complex conjugate of a derivative has been used earlier in [Kuz99] and has been called ‘smile conjugation’; we continue to use the same notation in this paper. Note that the smile conjugation simply treats and as local sections and applies the conjugation rule (2.62).
3 Hypermultiplets
The dynamical degrees of freedom of a hypermultiplet consists of two chiral superfields and such that form an doublet. The in question can be either or and the corresponding hypers are called standard and twisted hypermultiplets respectively. A standard hypermultiplet555See also [GR95] for a discussion in ordinary superspace. can be described in projective superspace either by an superfield [HL17] or by a pair of -arctic superfields . The analogous notation for the twisted hypers is and -arctic respectively. We describe free hypermultiplets in this section and study interactions in Section 5.
3.1 Standard hypermultiplets
3.1.1 superfield
We start with a complex superfield . In terms of the inhomogeneous coordinate , we have and
(3.1) |
The projective constraints give the following constraints on and :
(3.2) |
We see that and are chiral superfields since annihilates them. See Appendix A.2 for a review of superspace.
The superpartner fermions are defined as666The conjugate fermions are obtained as follows. is best thought of as which, under conjugation, goes to .
(3.3) |
The superpartners are in the doublet of ; they are also independent of since the above combinations are globally defined weight superfields, i.e., global holomorphic functions on which are indeed constants in . Using the expressions and that , we can arrive at the following superspace definitions for the :
(3.4) |
The next superfield in the multiplet would be which (1) is globally defined on , (2) has -weight , (3) is antisymmetric in , and (4) is a Lorentz vector. The only superfield which satisfies all these properties is , where . Thus, we have
(3.5) |
The above equations (3.5) can be explicitly checked by using the expressions for in (2.20), the complex conjugate derivatives in (2.70), and the projective constraints (3.2).
The supersymmetric action that describes the (free) hypermultiplet is
(3.6) |
Using the fact that the superspace Lagrangian is annihilated by , we can write it as an action in superspace as in (2.26). We get
(3.7) |
Next, we can obtain the component action by first performing the -integral, pushing in the derivatives and using the definitions (3.4) and that and are chiral superfields:
(3.8) |
(See Appendix A.2 for a derivation of the component action from the action in the second line in (3.1.1).) We can also obtain the same component action as above by pushing in the derivatives in (3.6), use the definitions (3.3) and (3.5), and finally perform the integral (see Appendix C.2).
The superfield can be described in ordinary superspace as well. Writing and , the projective constraints are equivalent to
(3.9) |
As noted in [HL17], in contrast to an superfield in projective superspace, the above constraints do not put the superfield on-shell. Only the antisymmetric part in of is non-zero and it gives the superpartner fermions defined in (3.3) (or equivalently (3.4)):
(3.10) |
Note that the scalars are in an -doublet whereas the fermions are in an -doublet.
Recall from (3.2) that and are annihilated by and . Thus, we can write down a manifestly supersymmetric action with the measure :
(3.11) |
This is the projective superspace action (3.6) after plugging in , and performing the integral; therefore, it also coincides with the action (3.7). The above action is not manifestly R-symmetric, but a manifestly R-symmetric action also exists which agrees with any of the above actions (up to total spacetime derivatives):
(3.12) |
However, the above action is not manifestly supersymmetric since the measure does not involve all four superspace derivatives.
The -projective superspace action (3.6) does not seem to be invariant under R-symmetry since the Lagrangian does not seem to transform with -weight . To write a manifestly R-symmetric action in projective superspace, we use the arctic realization of the hypermultiplet, one that arises naturally from projective superspace (see Appendix B.2).
3.1.2 Arctic superfield
Consider two arctic multiplets and with -expansions
(3.13) |
The projective constraints give
(3.14) |
and similarly for . The zeroth components and are chiral superfields since whereas the , , , are unconstrained as superfields.
The supersymmetric action that describes the (free) standard hypermultiplet is
(3.15) |
In fact, the above action is equivalent to that of an superfield when we go partially on-shell by performing the -integral in the last two terms and integrating out the fields for :
(3.16) |
Since the , , are unconstrained as superfields, we can integrate them out in the above superspace action. This imposes for and retains only the and terms in . Integrating out , we get for . We cannot integrate out in the same way and set since is constrained as a superfield, . Instead, integrating out the constrained superfield constrains to satisfy .777Here is the procedure to integrate out a constrained superfield: we first relax the constraint on and introduce a Lagrange multiplier superfield : . Integrating out re-imposes the constraint whereas integrating out gives , which indeed satisfies . We can conclude the same by going down to components, or at an intermediate stage by pushing in in the first term in the Lagrangian (3.1.2) to get . Since the remaining measure does not kill , we can integrate it out to conclude that .
Thus, we have two chiral superfields and which we relabel as and respectively to make contact with the superfield terminology (3.1). Thus, becomes an superfield when we go partially on-shell by integrating out the auxiliary superfield :
(3.17) |
and the action (3.15) becomes the action (3.1.1):
(3.18) |
Since integrating out gives an superfield, it is consistent to give an -weight of to . However, the action does not seem to have -weight and hence does not appear R-symmetric. But the action in ordinary space (3.18) is certainly R-symmetric! Let us see how to understand the R-symmetry of (3.15).
The terms depending on can be made to have weight by declaring that is a weight superfield. However, the kinetic term is still a problem. Since is an auxiliary superfield, we can give it a non-standard R-symmetry transformation so that it cancels that of the kinetic term (this is motivated from the reduction in Appendix B.2):
(3.19) |
Recall that we must only perform infinitesimal -transformations on arctic superfields (see the discussion above eq. (2.46)). It is easy to check that the Lagrangian (3.15) transforms with weight zero when we transform according to the above rule (see Appendix B.2 for an explicit demonstration).
We could write down the descendants directly by acting on and with the derivatives . We could then compute the component action (3.18) by pushing the derivatives in the measure into the Lagrangian in the action (3.15) and using the definition of the descendants. This procedure results in the same conclusions, namely that is truncated to an superfield and is auxiliary, and hence we do not describe it here. However, see Appendix C.2 for an illustration of this method for an arctic fermi superfield.
3.2 Twisted hypermultiplets
A twisted hypermultiplet is described by a complex superfield that is specified as
(3.20) |
The -projective constraints are given by
(3.21) |
and are chiral superfields since annihilates them. As for the standard hyper, the superpartner fermions are defined by
(3.22) |
The above definition makes it clear that the superpartner fermions of are in the doublet of . Explicitly, we have
(3.23) |
The supersymmetric action that describes the twisted hypermultiplet is
(3.24) |
Performing the integral, we get
(3.25) |
which is the action for two chiral multiplets and . The form an -doublet and hence, the above multiplet describes a twisted hyper. The description in terms of -arctic superfields is analogous to that of the standard hyper.
4 Fermi multiplets
In this section, we describe matter fermi multiplets. We focus on -projective fermi superfields below; the -case follows analogously. Like hypermultiplets, fermi multiplets can be realized either as superfields or -arctic superfields. We only describe arctic superfields here since all our constructions use only those and not the superfields.
Start with a weight -arctic superfield satisfying
(4.1) |
The constraints in terms of are
(4.2) |
The for are unconstrained superfields while satisfies the chirality constraint . We relabel as . The action is
(4.3) |
In the last step, we have integrated out the (0,2) unconstrained superfields with . Note that this is consistent with the -transformations discussed in Section 2.5 only for weight . In more detail, the transformation rules for the fields in (2.48) preserve the auxiliary field equations , , only for weight .
To get the component action, we push the measure derivatives into the Lagrangian:
(4.4) |
where the auxiliary field is defined as . In Appendix C.2, we define the ordinary space components of directly without going to superspace by acting on with successively. We also compute the above component action by directly pushing in the measure in (4) and using the definitions of the components that were just alluded to, and finally perform the -integral.
5 Interactions
The criteria for supersymmetry are closure of the algebra on all the superfields and the invariance of the action (see the comments at the end of Section 2.2). In this section, we use these criteria to discover possible supersymmetric interactions between twisted hypers, standard hypers and fermis.
As indicated in the Introduction (Section 1), interactions could be -terms, gauge interactions, or of the nonlinear sigma model type. Nonlinear sigma models have been discussed for standard hypers in projective superspace [HL17a] and arctic standard hypers in projective superspace [LR88]. We have not explored all the possibilities for -term interactions. In this paper, we consider the combination of -arctic standard hypers, -arctic fermis and twisted hypers with the R-charge assignments given previously (of course, everything we say can be used for the mirror combination where we swap the two R-symmetry groups).
Consider -arctic fermi multiplets , arctic standard hypermultiplets and twisted hypermultiplets with the following projective constraints:
(5.1) |
where is the fully contracted derivative (see (2.29) in Section 2), and are superfields which are functions of the various superfields in the model. The second line in (5) is obtained by applying extended complex conjugation on the first line and using the appropriate definitions of extended complex conjugates from Section 2.6.
The closure of the supersymmetry algebra on and give
(5.2) |
The action for the above superfields splits into an action in -projective superspace for the standard hypers and the fermis, and an action in -projective superspace for the twisted hypers, i.e., with
(5.3) |
The action and in (5) are invariant if the Lagrangians are annihilated by . This is obvious for . The action of on the Lagrangian in is
(5.4) |
For the right hand side to be zero, the following conditions then have to be satisfied:
(5.5) |
(the two conditions are consistent with each other since we have for an superfield .)
Upon using (5.5), the constraints in (5.2) become
(5.6) |
and are a priori functions of both standard and twisted hypers. We restrict ourselves to the case where and are polynomials in the standard and twisted hypers. Recall that the -weights of , , and are , , and respectively. Since the -weight has to be preserved in the constraint equations (5) above, and should have -weight . Further, since we restrict and to be polynomials in the superfields, they must simply be independent of the standard hypers .
The reality constraints (5.5) are also consistent with and being independent of standard hypers. However, note that and can be chosen to be more general -weight 0 functions of the standard hypers (e.g. rational functions) and these may have good Taylor expansions around both and . Then it is possible to satisfy the reality constraint (5.5) even when and depend on arctic standard hypers non-trivially.
Since is an superfield which is assumed to be a polynomial in the twisted hypers and is annihilated by , it must be linear in the twisted hypers . Thus, must take the form
(5.7) |
where is and constant, and is constant.
Recall from Sections 3.1, 3.2 and 4 that the dynamical components of the arctic standard hyper are , those of the twisted hyper are and that of the fermi is . The full component action for these fields that follows from the projective superspace action (5) is worked out in Appendix C. We give the result here:
(5.8) |
6 Example: ADHM sigma model
In this section we consider an interacting model with standard hypers, fermis and twisted hypers. This is a particular linear sigma model which flows to a nonlinear sigma model with target space a -instanton solution in Yang-Mills theory in four dimensions. This model was written in superspace in [Dou98, Wit95] and in harmonic superspace in [GS95, GS96].
This linear sigma model for instantons is realised by the following nested D-brane configuration in Type IIB theory [Dou98]: 1 D1-brane D5-branes D9-branes. The D5-branes appear as -instanton configurations in the D9-brane gauge theory and the D1-brane probes this configuration. The dimensional linear sigma model is the theory on the D1-brane worldsheet.
The D1-brane worldsheet theory includes a gauge multiplet arising from the D1-D1 open string spectrum. However, the multiplet does not have an effect on the computation of the instanton connection on target space in the classical theory on the D1-brane [Dou98]. We describe the classical instanton model without the gauge multiplet in Section 6.1 and show that it reproduces the calculation in [Dou98], and redo the analysis more carefully in the companion paper [PR] with the gauge multiplet included. The novelty of the projective superspace approach is that twistor space and the relevant holomorphic bundles on twistor space required for describing instantons [War77, AW77, AHDM78] appear explicitly in the description of the model which we describe below.
For instantons, we add an O9--plane to the above D-brane configuration. The orientifold projection requires an even number of D5-branes which we take to be , and after the projection pairs of D5-branes are stuck and cannot be separated. The projection reduces the D9-brane gauge group to , that of the D5-branes to and projects out the vector multiplet on the D1-brane. For instantons, we start with D9-branes, D5-branes and D1-branes and add an O9+-plane which results in an gauge group on the D9-branes, an gauge group on the D5-branes and an gauge group on the D1-branes (again, the two D1-branes cannot be separated). These facts may be found in, e.g., [GP96]. Since the instanton sigma model requires a gauge multiplet, and both and models require orientifolds, we describe both sigma models together in the companion paper [PR].
6.1 instantons
The superfield content consists of
-
1.
twisted hypers , (we consider for most of the discussion),
-
2.
standard hypers , ,
-
3.
fermis , .
The above superfields (for ) are a subset of the low-energy spectrum of the various D-D open strings in the D-brane configuration described above. Since we are interested in the low-energy theory on the D1-brane, we retain only those fields that appear from the D1-D open string sectors for . The two twisted hypers arise from the D1-D1 strings in the directions transverse to the D1-brane and D5-branes. The standard hypers arise from D1-D5 strings and the fermis arise from the D1-D5 strings ( fermis) and the D1-D9 strings ( fermis). Part of the couplings described below arise from the D5-D9 open string degrees of freedom which are frozen from the point of view of the D1-brane, and they contain the instanton moduli.
We suppress the flavour indices , and on the twisted hypers, standard hypers and fermis respectively unless we wish to explicitly exhibit the flavour properties of the superfields. We work with a given symplectic structure on the space of twisted hypers. This allows for a reality condition:
(6.1) |
(note that according to the above condition since . This is consistent with the result for an multiplet). The most general constraints are those given in (5):
(6.2) |
where recall from Section 5 that , are superfields. As discussed in Section 5, and are independent of the standard hypers and are linear in the twisted hypers . The constraints on the couplings and that follow from the closure of the superalgebra are (5.2) which we reproduce here for convenience:
(6.3) |
and and are and matrices respectively; with the flavour indices explicitly displayed, the matrices are resp. written as and . Recall from the discussion around (5.7) that has to be of the form
(6.4) |
where are the twisted hypermultiplets. The coupling is a constant matrix superfield and the are constant matrices (one matrix for each ).
Twistor space
Let us consider two twisted hypers, i.e., (everything we say for two twisted hypers can be extended to general ). The twisted hyper superfields are coordinates on the target space . The doublet together with the projective superfields can be interpreted as homogeneous coordinates for a which is in fact the twistor space of (the one-point compactification of the target ). The symplectic structure on the space of twisted hypers and the symplectic structure on the space of -doublets together give an antiholomorphic involution , , on the which squares to . The serve as coordinates on the correspondence space and the incidence relations are simply the definition of the as projective superfields.
Monads on twistor space
Next, we show that the couplings and encode the data of a monad on . Let and be the vector spaces of and respectively with and be the vector space of fermis with . Then, the couplings and can be interpreted as elements of and respectively, as is clear from the constraints (6.2). Recall that these maps are linear in the homogeneous coordinates since . We thus have
(6.5) |
The constraint that follows from the closure of the algebra (6.3) makes (6.5) a complex. We further require that is injective and is surjective: this imposes non-degeneracy conditions on the couplings and . Then the above complex is precisely a monad and the cohomology at , i.e., is a holomorphic rank vector bundle on which is trivial when restricted to lines in , and has . Thus, the data that goes into choosing the off-shell superfield content of our linear sigma model is precisely the same data that goes into defining a holomorphic bundle on twistor space that is trivial on lines.
We get a symplectic structure on the bundle also from the requirement that the action is supersymmetric. Some reality conditions (which were implicit in the previous sections) are necessary on the vector spaces , and to write down an action for the projective superfields , and . They are (1) a hermitian structure on that identifies , and (2) the identification , where and stand for the complex conjugate and dual of a vector space respectively. With these at hand, the invariance of the action gives the following constraint (5.5) on the couplings and :
(6.6) |
where the bar on acts the hermitian conjugate on the matrix components and extended conjugate on the superfield. This imposes a symplectic structure on the bundle obtained from the monad (6.5). By the Penrose-Ward-Atiyah correspondence [War77, AW77], the bundle on twistor space with the symplectic structure described above corresponds to a self-dual connection on (more precisely, on the one-point compactification of ). The ADHM construction [AHDM78] gives an explicit expression for the instanton gauge field in terms of the data described above. The constraints are precisely the ADHM equations that describe the instanton moduli space [AHDM78].
Next, we show that the model flows to an instanton solution in the infrared by explicitly obtaining the expression for the instanton gauge field given by the ADHM construction [AHDM78]. The material in the rest of this section is not new and follows the calculations in [Wit95, Dou98]. In Section 6.2 below, we choose particular bases for the vector spaces of superfields to give the usual standard characterization of the ADHM instanton moduli space in terms of finite dimensional matrices. Again, most of the material is standard except for a formula of the virtual dimension of the instanton moduli space on for .
Instantons on
The potential energy density of the model described above can be read off from the general expression in (5) and is positive-definite:
(6.7) |
Recall that and the are components of the arctic standard hyper once we eliminate the auxiliary superfields accompanying higher powers of (see (3.17) and the discussion around it). Suppose the constant matrices and are sufficiently generic so that is an invertible matrix, i.e., all its eigenvalues are non-zero, for any value of . Then, the vacuum corresponds to setting the for every flavour .
About this vacuum, the potential vanishes and in particular does not give a mass for the twisted hyper scalars: there is a classical moduli space of vacua parametrized by the four twisted hyper scalars with the reality condition (6.1). Under the genericity assumption on and , the eigenvalues of the standard hyper mass matrix are all (1) positive since is a positive-definite matrix, and (2) strictly positive since is invertible. We list them as . Then, the mass of the standard hyper scalars for a given is . The Yukawa couplings can also be read off from (5):
(6.8) |
On the classical vacuum moduli space characterised by and arbitrary , the twisted hyper fermions again have no mass terms. Let us look at the mass terms for the standard hyper fermions :
(6.9) |
where we have displayed the flavour indices explicitly. Recall that we have diagonalized . By using an appropriate transformation, we can further cast the matrix into a block form with a non-trivial block and a zero block:
(6.10) |
where the non-trivial block is . For a fixed flavour of the standard hyper, the two fermions , and the two fermis , interact through the mass matrix and become massive with mass . Recall that the standard hyper scalars also have the same mass . The zero block of size implies that the fermis , are massless. Thus, for generic values of the couplings and , we have massive standard hypers, massive fermis and massless fermis about any point of the classical vacuum moduli space that is parametrized by the massless twisted hypers.
The massless fermis can be characterised more generally as the solutions of the equation
(6.11) |
Let the massless solutions be arranged into the matrix with the normalisation . The most general massless solution is then
(6.12) |
Plugging in the above expression for in its kinetic term, we get the kinetic term for the massless modes :
(6.13) |
We see that the massless fermis have now acquired an additional connection which is the pullback of a connection on target space :
(6.14) |
This is the connection for a -instanton solution with gauge group, a fact that follows from standard results in the ADHM construction. Since we have assumed the instanton to be non-degenerate, the part of the connection is trivial and is in fact an instanton connection. We study the degenerate cases carefully in [PR] where we shall find that the gauge multiplet on the D1-brane worldsheet plays an important role.
6.2 The instanton moduli space and symmetries
The constraints and the fermi zero modes (6.11) (and in turn, the formula for the instanton gauge field) are unaffected by transformations on the space of standard hypermultiplets and transformations of the space of fermis:
(6.15) |
Thus, two different solutions of that are related by a transformation as in (6.15) correspond to the same instanton solution. This redundancy allows us to choose a simple form for the coupling and the equations .
Plugging in the explicit form , we get
(6.16) |
We have used the reality condition (6.1) on the twisted hypers in going to the second line above. Terms with different numbers of twisted hypers must vanish separately. Let us study each of them in turn:
-
1.
The constant part of (6.2) satisfies . Displaying the indices explicitly, we have
(6.17) where is a positive-definite matrix.
-
2.
The vanishing of the terms linear in in (6.2) requires
(6.18) -
3.
The term quadratic in the twisted hypers vanishes when
(6.19) where is antisymmetric in and is an arbitrary hermitian matrix for each . For the special case , i.e., when there are two twisted hypers, the antisymmetric matrix is proportional to the symplectic form :
(6.20) where is now a single positive-definite matrix.
The couplings and transform under the (6.15) as
(6.21) |
with the same matrix and matrix for all . This freedom can be used to choose a convenient form for and as follows.
First, the satisfy the constraints (6.19) . Suppose we choose the symplectic form canonically to be
(6.22) |
where indicates that length of the diagonal matrix is . Let us look at the pair of matrices , . They satisfy
(6.23) |
By an appropriate transformation (6.21), we can transform into the identity matrix. Then, the matrix satisfies
(6.24) |
where and are the identity and zero matrices respectively. Using an appropriate transformation (6.21), we can cast the above matrix into the form
(6.25) |
There is a residual subgroup of which preserves the above configuration (6.25) which corresponds to
(6.26) |
The reality constraint (6.18) for , i.e.,
(6.27) |
is solved by the following expression for :
(6.28) |
where , are matrices and , are matrices. The remaining matrices , can also be simplified to a form similar to (6.28) using the constraints
(6.29) |
where we have introduced the index , such that the pairs cover the index (later, we will append the value as well). We then get the simplified form
(6.30) |
Thus, the degrees of freedom that remain after fixing the symmetries are
(6.31) |
There real degrees of freedom.
The remaining constraints on the and , , are
(6.32) |
In terms of the matrices , , and , , we have the equations
(6.33) |
Let us get a count of the number of such equations. The above equations are symmetric in , . For , the last equation in (6.2) is manifestly real whereas the first equation is complex. Thus, for , we have real equations. For , it is sufficient to restrict , and both equations in (6.2) are complex. This gives a count of . In total, the number of equations is . For , the target space is and the above equations are precisely the ADHM equations.
We must also remember that the instanton connection (6.14) is invariant under the residual transformations (6.26). We treat the residual in (6.26) as a symmetry of framings at of the instanton solution. The , are inert under framing whereas the and transform as
(6.34) |
Thus, the moduli space of framed instantons is described by
(6.35) |
with
-
1.
Fields: , ,
-
2.
Equations: the equations (6.2), and
-
3.
Symmetries: the residual symmetry in (6.26) which acts on the various fields as
(6.36)
The virtual dimension of the moduli space of framed instantons is then
(6.37) |
When , this becomes which is the virtual dimension (in fact, the dimension itself) of the -instanton moduli space on .
Acknowledgements
The authors thank N. Nekrasov,
S. Shatashvili and J. P. Ang for useful discussions during early
stages of the work. N. P. would like to acknowledge the support of the
Jawaharlal Nehru Postdoctoral Fellowship for part of the duration of
the work. The work of M. R. was supported in part by NSF grants
PHY-19-15093 and PHY-22-15093.
Appendix A and superspace
A.1 superspace
superspace has coordinates where is a real Grassmann variable. The corresponding supercovariant derivatives are which satisfy the algebra
(A.1) |
with all other commutators being zero.
Multiplets of the supersymmetry algebra are not constrained. The most common ones are the scalar multiplet (spin ), the fermi multiplet (spin , left-handed) and the gauge multiplet (spin ). The multiplets are irreducible representations of the algebra when they are real (or hermitian).
A real scalar superfield has components
(A.2) |
where is a real scalar field and is a real right-handed fermion. We follow the usual convention of denoting the lowest component of a superfield by the same symbol and drop the ‘slash’ from here on. A supersymmetric action with the lowest number of derivatives is
(A.3) |
A real fermi superfield has the components
(A.4) |
where is a real left-handed fermion and is a real auxiliary field, with the action
(A.5) |
One can add a potential term in the action via a term that is linear in the fermi superfields in the theory:
(A.6) |
where are functions of the scalar superfields in the theory.
A.2 superspace
superspace has coordinates where and are left-handed spinors. We denote the corresponding supercovariant derivatives by . They satisfy the algebra
(A.7) |
We review various constrained superfields that are required to write down supersymmetric actions in superspace.
Chiral
A scalar chiral superfield (or, simply a chiral superfield) is a Lorentz scalar and satisfies and has components
(A.8) |
and consequently, . The action for a free chiral superfield is
(A.9) |
Fermi
A Fermi superfield is a left-handed spinor and satisfies the constraint . It has components
(A.10) |
The action for a free Fermi multiplet is
(A.11) |
We see that the left-handed fermion satisfies the equation of motion and hence is right-moving on-shell. The field is auxiliary with equation of motion .
Potential terms
Let collectively denote all the chiral superfields in the theory and the Fermi superfields. We can modify the constraint to
(A.12) |
where the are holomorphic functions of the chiral multiplets . This modification results in additional interaction terms in the action for the fermi superfields:
(A.13) |
We can also write a superpotential term, known as a “-term” in literature:
(A.14) |
Since the superspace measure in the -term involves only half the supercovariant derivatives, its invariance under supersymmetry requires the integrand to be chiral, i.e., . This implies
(A.15) |
If the above constraint is not satisfied, supersymmetry is softly broken down from to , even though the -term is written in superspace.
Reduction to superspace
Define the derivatives
(A.16) |
is the real super derivative and is the generator of the extra (non-manifest) supersymmetry.
The chiral and fermi multiplets (and their antichiral counterparts) become complex scalar and fermi multiplets with components
(A.17) |
We have . We can discard the second term since it gives rise to a total derivative term in the action. Using that acts as on superfields satisfying , we can write the actions in superspace:
(A.18) |
where .
Appendix B projective superspace and
B.1 Definitions
We start with the real supercharges with , . The R-symmetry group is
We restrict our attention to the subgroup where , the diagonal subgroup of , and . The supercharges can then be written as where , and are doublet indices of , and respectively. This restriction of the R-symmetry group to a subgroup seems to be required to obtain the vector multiplet via gauged supercovariant derivatives and the relevant superspace constraints [Sie85].
The algebra of supercovariant derivatives and is
(B.1) |
The reality conditions on the derivatives are
(B.2) |
It will be useful to define the subalgebra spanned by the derivatives
(B.3) |
which satisfy
(B.4) |
The non-manifest supersymmetry generators are then , and , .
The general projective superspace corresponding to is described by introducing a doublet for each of the s in the R-symmetry group: , and for the subgroups , and respectively.
We then define the following projective supercovariant derivatives:
(B.5) |
We also introduce the doublets , and as was done in the main text above eq. (2.9). We again choose , and and define the linearly independent derivatives:
(B.6) |
We consider projective superfields which are functions of one projective coordinate from the left moving sector ( or ) and one projective coordinate from the right-moving sector ( or ) and are annihilated by the corresponding set of projective derivatives. For example, an projective superfield is a function of and and is annihilated by and . The supersymmetric action is
(B.7) |
Using that is annihilated by and and , , we can replace the measure by the measure and do the , integrals to get an action in superspace:
(B.8) |
There are many choices for projective superfields: they can be a polynomial or a power series in each of the projective coordinates that they depend on. A polynomial superfield with respect to , and will be respectively denoted as , and . Power series superfields are typically denoted as -arctic, -antarctic, -arctic and so on. Below, we discuss the arctic superfield, i.e., an arctic superfield which is a power series only in and is annihilated by and .
B.2 standard hypermultiplet
Consider an arctic superfield with alternate notation and for and respectively. The constraints give the constraints
(B.9) |
is chiral as an superfield since , is complex linear since , whereas the are unconstrained as superfields.
The action for the arctic superfield is
(B.10) |
This action is R-symmetric since the measure has -weight ( from , from and from ) and the Lagrangian has -weight ( each from and , see the paragraph after equation (2.47) in Section 2.5).
Next, we obtain the content by applying to :
(B.11) |
Recall from (2.80) that the conjugate of when acting on arctic superfields is
(B.12) |
Using this, we get
(B.13) |
Using and , we get
(B.14) |
The supersymmetric action is obtained by pushing in the derivatives in the measure:
(B.15) |
Let us study the R-symmetry invariance of the above action in more detail. Recall that and (see (2.80)) transform under as
(B.16) |
where is a new superfield which is evaluated at whose expression is given by expanding the right hand side around . The transformations of all the other superfields can be obtained by using the above. We first summarize the results and then detailed calculations. The hypers , transform as weight objects:
(B.17) |
the fermis , transform as weight objects:
(B.18) |
and , transform as weight objects, along with an additional shift:
(B.19) |
Using these, the supersymmetric action (B.2) can be checked to be R-symmetric, a fact which was already demonstrated for the action (B.10).
The derivation of R-symmetry transformations
Note: In the following calculations, a ′ on superfields denotes the transformed superfield and must not be confused with the ′ on the R-symmetry indices.
Given the transformation of in (B.16), the transformation of is
(B.20) |
which implies that
(B.21) |
This tells us that transforms as a weight field as well. transforms as
(B.22) |
that is,
(B.23) |
Thus, transforms as a weight superfield but with an additional shift term proportional to . Finally, we need the transformation of . Analogous to (B.20), we have
(B.24) |
which gives
(B.25) |
Again, we see that transforms as a weight superfield, along with an additional shift term proportional to . We can also start with the definition of in (B.14) and arrive at the above result. In detail, we have
(B.26) |
Also using the transformation of from (B.21), we get
(B.27) |
The quantity in the parentheses in the last line simplifies to give . Plugging this into (B.2) and dividing by , we get (B.25).
Appendix C Component actions
In this appendix, we derive the action for the ordinary space components of the various superfields in two ways: (1) by reducing to superspace and using standard results from Appendix A.2, and (2) by reducing directly to ordinary space by pushing in the in the superspace measure.
C.1
Recall from Sections 3 and 4 the and expansions of the various superfields:
(C.1) |
Also recall that we relabelled some low-lying components of the above superfields since they were constrained as superfields:
(C.2) |
We reproduce here the projective superspace constraints on the various superfields given in (5):
(C.3) |
The actions are given by
(C.4) |
The closure of the projective superspace algebra on gives the constraints
(C.5) |
and the invariance of the above actions gives
(C.6) |
The assumption that the are polynomials in the various superfields constrains to take the form . The constraints (C.1) lead to the following -terms for the superfield and :
(C.7) |
Integrating out the auxiliary superfield proceeds in the same way as in the free case, with one important difference due to the -term for above. Unconstraining in the standard way (see Footnote 7), we get
(C.8) |
Integrating out gives which implies that is a chiral superfield which we labelled as . In addition, there is now a -term:
(C.9) |
where, in the last equality, we have used the constraint that follows from (C.5).
Rewriting the projective superspace measure in (C.1) as and performing the - and -integrals, we get the following superspace actions:
(C.10) |
with .
Now we further push in the derivatives in the actions above and compute the component actions according to Appendix A.2. Recall that the superspace components of are
(C.11) |
and the superspace components of are
(C.12) |
The components of the fermi are
(C.13) |
Let us work out the twisted hyper part of first. We have
(C.14) |
The standard hyper part of is given by
(C.15) |
whereas the fermi part of is given by
(C.16) |
Integrating out the auxiliary fields , , we get
(C.17) |
and the fermi action becomes
(C.18) |
Let us look at the potential terms:
(C.19) |
where, in the second step, we have used the reality constraint (C.6). The above form does not seem invariant under R-symmetry. However, it follows from that which allows us to write the potential in manifest R-symmetry form:
(C.20) |
Next, let us collect all the Yukawa couplings from the fermi action (C.1):
(C.21) |
We have . Then, we get
(C.22) |
We thus get the manifest R-symmetric form of the Yukawa couplings
(C.23) |
where the first term and its complex conjugate together are mass terms which contain the fermis and the superpartners of the standard hypers. The other terms are Yukawa couplings which involve the standard hypers, the twisted hypers and the fermis.
C.2
In this subsection, we directly go from superspace to ordinary space. We give two illustrative examples, an standard hyper and an arctic fermi.
standard hyper
Recall from (3.3) and (3.5) that the descendants of are, at the first level,
(C.24) |
and at the second level,
(C.25) |
The action is
(C.26) |
Pushing in the derivatives in the measure and using (C.24) and (C.25), we get
(C.27) |
where, in going to the second line, we have used the explicit expressions and (to compute , we follow the same steps as for in Section 2.6).
The arctic fermi superfield
We look at the weight arctic fermi superfield. The descendants are
(C.28) |
The action is
(C.29) |
Pushing in the derivatives in the measure, we get
(C.30) |
The superfields and are auxiliary and can be integrated out. The terms involving are
(C.31) |
Integrating out gives
(C.32) |
Thus, the weight superfield which was locally defined on becomes a constant on , which is nothing but a globally defined weight superfield. Integrating out just sets them to zero. Relabelling , the action becomes
(C.33) |
References
- [AHDM78] M. F. Atiyah, Nigel J. Hitchin, V. G. Drinfeld and Yu. I. Manin “Construction of Instantons” In Phys. Lett. A 65, 1978, pp. 185–187 DOI: 10.1016/0375-9601(78)90141-X
- [AKL07] Masato Arai, Sergei M. Kuzenko and Ulf Lindstrom “Hyperkahler sigma models on cotangent bundles of Hermitian symmetric spaces using projective superspace” In JHEP 02, 2007, pp. 100 DOI: 10.1088/1126-6708/2007/02/100
- [AW77] M. F. Atiyah and R. S. Ward “Instantons and Algebraic Geometry” In Commun. Math. Phys. 55, 1977, pp. 117–124 DOI: 10.1007/BF01626514
- [Ade+76] M. Ademollo “Supersymmetric Strings and Color Confinement” In Phys. Lett. B 62, 1976, pp. 105–110 DOI: 10.1016/0370-2693(76)90061-7
- [BD88] Tom Banks and Lance J. Dixon “Constraints on String Vacua with Space-Time Supersymmetry” In Nucl. Phys. B 307, 1988, pp. 93–108 DOI: 10.1016/0550-3213(88)90523-8
- [BDL96] Micha Berkooz, Michael R. Douglas and Robert G. Leigh “Branes intersecting at angles” In Nucl. Phys. B 480, 1996, pp. 265–278 DOI: 10.1016/S0550-3213(96)00452-X
- [BMG86] R. Brooks, F. Muhammad and S. J. Gates “Unidexterous D=2 Supersymmetry in Superspace” In Nucl. Phys. B 268, 1986, pp. 599–620 DOI: 10.1016/0550-3213(86)90261-0
- [BW76] L. Brink and J. O. Winnberg “The Superoperator Formalism of the Neveu-Schwarz-Ramond Model” In Nucl. Phys. B 103, 1976, pp. 445–464 DOI: 10.1016/0550-3213(76)90509-5
- [CGK85] Edward Corrigan, P. Goddard and A. Kent “Some Comments on the ADHM Construction in 4k-dimensions” In Commun. Math. Phys. 100, 1985, pp. 1 DOI: 10.1007/BF01212684
- [CHS91] Curtis G. Callan, Jeffrey A. Harvey and Andrew Strominger “World sheet approach to heterotic instantons and solitons” In Nucl. Phys. B 359, 1991, pp. 611–634 DOI: 10.1016/0550-3213(91)90074-8
- [CHS91a] Curtis G. Callan, Jeffrey A. Harvey and Andrew Strominger “Worldbrane actions for string solitons” In Nucl. Phys. B 367, 1991, pp. 60–82 DOI: 10.1016/0550-3213(91)90041-U
- [CHSW85] P. Candelas, Gary T. Horowitz, Andrew Strominger and Edward Witten “Vacuum Configurations for Superstrings” In Nucl. Phys. B 258, 1985, pp. 46–74 DOI: 10.1016/0550-3213(85)90602-9
- [DK94] Jacques Distler and Shamit Kachru “(0,2) Landau-Ginzburg theory” In Nucl. Phys. B 413, 1994, pp. 213–243 DOI: 10.1016/0550-3213(94)90619-X
- [DM96] Michael R. Douglas and Gregory W. Moore “D-branes, quivers, and ALE instantons”, 1996 arXiv:hep-th/9603167
- [DS86] Michael Dine and Nathan Seiberg “(2,0) SUPERSPACE” In Phys. Lett. B 180, 1986, pp. 364–369 DOI: 10.1016/0370-2693(86)91203-7
- [Dou98] Michael R. Douglas “Gauge fields and D-branes” In J. Geom. Phys. 28, 1998, pp. 255–262 DOI: 10.1016/S0393-0440(97)00024-7
- [EOTY89] Tohru Eguchi, Hirosi Ooguri, Anne Taormina and Sung-Kil Yang “Superconformal Algebras and String Compactification on Manifolds with SU(N) Holonomy” In Nucl. Phys. B 315, 1989, pp. 193–221 DOI: 10.1016/0550-3213(89)90454-9
- [FM73] D. B. Fairlie and D. Martin “New light on the Neveu-Schwarz model” In Nuovo Cim. A 18, 1973, pp. 373–383 DOI: 10.1007/BF02722834
- [GGP14] Abhijit Gadde, Sergei Gukov and Pavel Putrov “Duality Defects”, 2014 arXiv:1404.2929 [hep-th]
- [GGP16] Abhijit Gadde, Sergei Gukov and Pavel Putrov “Fivebranes and 4-manifolds” In Prog. Math. 319, 2016, pp. 155–245 DOI: 10.1007/978-3-319-43648-7˙7
- [GGRS83] S. J. Gates, Marcus T. Grisaru, M. Rocek and W. Siegel “Superspace Or One Thousand and One Lessons in Supersymmetry” 58, Frontiers in Physics Benjamin-Cummings Publishing Company, 1983 arXiv:hep-th/0108200
- [GHM97] Michael B. Green, Jeffrey A. Harvey and Gregory W. Moore “I-brane inflow and anomalous couplings on D-branes” In Class. Quant. Grav. 14, 1997, pp. 47–52 DOI: 10.1088/0264-9381/14/1/008
- [GHMR85] David J. Gross, Jeffrey A. Harvey, Emil J. Martinec and Ryan Rohm “The Heterotic String” In Phys. Rev. Lett. 54, 1985, pp. 502–505 DOI: 10.1103/PhysRevLett.54.502
- [GHR84] S. J. Gates, C. M. Hull and M. Rocek “Twisted Multiplets and New Supersymmetric Nonlinear Sigma Models” In Nucl. Phys. B 248, 1984, pp. 157–186 DOI: 10.1016/0550-3213(84)90592-3
- [GIKO+84] A. Galperin et al. “Unconstrained N=2 Matter, Yang-Mills and Supergravity Theories in Harmonic Superspace” [Erratum: Class.Quant.Grav. 2, 127 (1985)] In Class. Quant. Grav. 1, 1984, pp. 469–498 DOI: 10.1088/0264-9381/1/5/004
- [GIOS84] A. Galperin, E. Ivanov, V. Ogievetsky and E. Sokatchev “Harmonic Superspace: key to N=2 Supersymmetry Theories” In JETP Lett. 40, 1984, pp. 912–916
- [GK99] S. James Gates and Sergei M. Kuzenko “The CNM hypermultiplet nexus” In Nucl. Phys. B 543, 1999, pp. 122–140 DOI: 10.1016/S0550-3213(98)00870-0
- [GL85] Johan Grundberg and Ulf Lindstrom “Actions for Linear Multiplets in Six-dimensions” In Class. Quant. Grav. 2, 1985, pp. L33 DOI: 10.1088/0264-9381/2/2/005
- [GMMS05] Sergei Gukov, Emil Martinec, Gregory W. Moore and Andrew Strominger “The Search for a holographic dual to AdS(3) x S**3 x S**3 x S**1” In Adv. Theor. Math. Phys. 9, 2005, pp. 435–525 DOI: 10.4310/ATMP.2005.v9.n3.a3
- [GP96] Eric G. Gimon and Joseph Polchinski “Consistency conditions for orientifolds and D-manifolds” In Phys. Rev. D 54, 1996, pp. 1667–1676 DOI: 10.1103/PhysRevD.54.1667
- [GR95] S. James Gates and Lubna Rana “Manifest (4,0) supersymmetry, sigma models and the ADHM instanton construction” In Phys. Lett. B 345, 1995, pp. 233–241 DOI: 10.1016/0370-2693(94)01653-T
- [GS95] A. Galperin and E. Sokatchev “Manifest supersymmetry and the ADHM construction of instantons” In Nucl. Phys. B 452, 1995, pp. 431–455 DOI: 10.1016/0550-3213(95)00272-T
- [GS96] A. Galperin and E. Sokatchev “Supersymmetric sigma models and the ’t Hooft instantons” In Class. Quant. Grav. 13, 1996, pp. 161–170 DOI: 10.1088/0264-9381/13/2/004
- [HKLR87] Nigel J. Hitchin, A. Karlhede, U. Lindstrom and M. Rocek “Hyperkahler Metrics and Supersymmetry” In Commun. Math. Phys. 108, 1987, pp. 535 DOI: 10.1007/BF01214418
- [HL17] Chris Hull and Ulf Lindström “All : Sigma Models with Off-Shell Supersymmetry” In JHEP 08, 2017, pp. 129 DOI: 10.1007/JHEP08(2017)129
- [HL17a] Chris Hull and Ulf Lindström “All : Sigma Models with Off-Shell Supersymmetry” In JHEP 03, 2017, pp. 042 DOI: 10.1007/JHEP03(2017)042
- [HW85] C. M. Hull and Edward Witten “Supersymmetric Sigma Models and the Heterotic String” In Phys. Lett. B 160, 1985, pp. 398–402 DOI: 10.1016/0370-2693(85)90008-5
- [IKS06] Nissan Itzhaki, David Kutasov and Nathan Seiberg “I-brane dynamics” In JHEP 01, 2006, pp. 119 DOI: 10.1088/1126-6708/2006/01/119
- [IR96] I. T. Ivanov and M. Rocek “Supersymmetric sigma models, twistors, and the Atiyah-Hitchin metric” In Commun. Math. Phys. 182, 1996, pp. 291–302 DOI: 10.1007/BF02517891
- [JS09] Dharmesh Jain and Warren Siegel “Deriving Projective Hyperspace from Harmonic” In Phys. Rev. D 80, 2009, pp. 045024 DOI: 10.1103/PhysRevD.80.045024
- [KLR84] Anders Karlhede, Ulf Lindstrom and Martin Rocek “Selfinteracting Tensor Multiplets in Superspace” In Phys. Lett. B 147, 1984, pp. 297–300 DOI: 10.1016/0370-2693(84)90120-5
- [KLR87] A. Karlhede, U. Lindstrom and M. Rocek “Hyperkahler Manifolds and Nonlinear Supermultiplets” In Commun. Math. Phys. 108, 1987, pp. 529 DOI: 10.1007/BF01214417
- [Kuz99] Sergei M. Kuzenko “Projective superspace as a double punctured harmonic superspace” In Int. J. Mod. Phys. A 14, 1999, pp. 1737–1758 DOI: 10.1142/S0217751X99000889
- [LR10] Ulf Lindstrom and Martin Rocek “Properties of hyperkahler manifolds and their twistor spaces” In Commun. Math. Phys. 293, 2010, pp. 257–278 DOI: 10.1007/s00220-009-0923-0
- [LR83] U. Lindstrom and M. Rocek “Scalar Tensor Duality and N=1, N=2 Nonlinear Sigma Models” In Nucl. Phys. B 222, 1983, pp. 285–308 DOI: 10.1016/0550-3213(83)90638-7
- [LR88] U. Lindstrom and M. Rocek “New Hyperkahler Metrics and New Supermultiplets” In Commun. Math. Phys. 115, 1988, pp. 21 DOI: 10.1007/BF01238851
- [LR90] U. Lindstrom and M. Rocek “ Superyang-mills Theory in Projective Superspace” In Commun. Math. Phys. 128, 1990, pp. 191 DOI: 10.1007/BF02097052
- [MSW97] Juan Martin Maldacena, Andrew Strominger and Edward Witten “Black hole entropy in M theory” In JHEP 12, 1997, pp. 002 DOI: 10.1088/1126-6708/1997/12/002
- [NP17] Nikita Nekrasov and Naveen S. Prabhakar “Spiked Instantons from Intersecting D-branes” In Nucl. Phys. B 914, 2017, pp. 257–300 DOI: 10.1016/j.nuclphysb.2016.11.014
- [Nek16] Nikita Nekrasov “BPS/CFT correspondence: non-perturbative Dyson-Schwinger equations and qq-characters” In JHEP 03, 2016, pp. 181 DOI: 10.1007/JHEP03(2016)181
- [PR] Naveen S. Prabhakar and Martin Roček “N=(0,4) projective superspace: gauge multiplets” in preparation
- [PSY16] Pavel Putrov, Jaewon Song and Wenbin Yan “(0,4) dualities” In JHEP 03, 2016, pp. 185 DOI: 10.1007/JHEP03(2016)185
- [RS86] A. A. Roslyi and Albert S. Schwarz “Supersymmetry in a space with auxiliary dimensions” In Commun. Math. Phys. 105, 1986, pp. 645 DOI: 10.1007/BF01238937
- [Ros83] AA Rosly “Super Yang-Mills constraints as integrability conditions” In Proceedings of the International Seminar on Group Theoretical Methods in Physics,(Zvenigorod, USSR, 1982), MA Markov (Ed.), Nauka, Moscow 1, 1983, pp. 263
- [Ros85] A A Rosly “Gauge fields in superspace and twistors” In Classical and Quantum Gravity 2.5, 1985, pp. 693 DOI: 10.1088/0264-9381/2/5/011
- [Sak85] Makoto Sakamoto “/2 Supersymmetry in Two-dimensions” In Phys. Lett. B 151, 1985, pp. 115–118 DOI: 10.1016/0370-2693(85)91396-6
- [Sei88] Nathan Seiberg “Observations on the Moduli Space of Superconformal Field Theories” In Nucl. Phys. B 303, 1988, pp. 286–304 DOI: 10.1016/0550-3213(88)90183-6
- [Sie84] W. Siegel “Manifest Lorentz Invariance Sometimes Requires Nonlinearity” In Nucl. Phys. B 238, 1984, pp. 307–316 DOI: 10.1016/0550-3213(84)90453-X
- [Sie85] W. Siegel “Some extended supersymmetric two-dimensional scalar multiplets” In Class. Quant. Grav. 2, 1985, pp. L41 DOI: 10.1088/0264-9381/2/3/001
- [Ton14] David Tong “The holographic dual of ” In JHEP 04, 2014, pp. 193 DOI: 10.1007/JHEP04(2014)193
- [War77] R. S. Ward “On Selfdual gauge fields” In Phys. Lett. A 61, 1977, pp. 81–82 DOI: 10.1016/0375-9601(77)90842-8
- [Wit95] Edward Witten “Sigma models and the ADHM construction of instantons” In J. Geom. Phys. 15, 1995, pp. 215–226 DOI: 10.1016/0393-0440(94)00047-8
- [Zum75] Bruno Zumino “Supersymmetry” In Conf. Proc. C 750926, 1975, pp. 255–280