Notes on Derived Deformation Theory for Field Theories and Their Symmetries
Abstract
1. Introduction
Before discussing details, I want to say clearly at the outset: the wonderful machine of modern algebraic geometry was created to understand basic and naive questions about geometry (broadly construed). The purpose of this book is to give you a thorough foundation in these powerful ideas. Do not be seduced by the lotus-eaters into infatuation with untethered abstraction. Hold tight to your geometric motivation as you learn the formal structures which have proved to be so effective in studying fundamental questions. When introduced to a new idea, always ask why you should care. Do not expect an answer right away, but demand an answer eventually. Try at least to apply any new abstraction to some concrete example you can understand well. See if you can make a rough picture to capture the essence of the idea. (I deliberately asked an uncoordinated and confused three-year-old to make most of the figures in the book in order to show that even quick sketches can enlighten and clarify.)
Understanding algebraic geometry is often thought to be hard because it consists of large complicated pieces of machinery. In fact the opposite is true; to switch metaphors, rather than being narrow and deep, algebraic geometry is shallow but extremely broad. It is built out of a large number of very small parts, in keeping with Grothendieck’s vision of mathematics. It is a challenge to hold the entire organic structure, with its messy interconnections, in your head.
Those who have meditated on the beauty and utility, in theoretical mechanics, of the general method of Lagrange—who have felt the power and dignity of that central dynamical theorem which he deduced, in the Mécanique Analytique, from a combination of the principle of virtual velocities with the principle of d’Alembert—and who have appreciated the simplicity and harmony which he introduced into the research of the planetary perturbations, by the idea of the variation of parameters, and the differentials of the disturbing function, must feel that mathematical optics can only then attain a coordinate rank with mathematical mechanics, or with dynamical astronomy, in beauty, power, and harmony, when it shall possess an appropriate method, and become the unfolding of a central idea.
1.1. How to Read This Article
- §1:
- Introduction. Self-explanatory, and hopefully also explanatory in a broader way.
- §2:
- Physicists should have invented sheaves. By thinking about the principle of general covariance, we arrive (in a massively simplified setting) at the notion of an “étale site:” a category of structured manifolds, related by maps that are locally structure-preserving isomorphisms. By thinking about local fields, we arrive at the notion of a sheaf on this site.
- §3:
- Subtle objects are best built from boring ones. By using a system of local observers to reconstruct global information, we rediscover Čech cohomology. Thinking about local equations of motion as a source of propagation of information, we rediscover de Rham cohomology and the equivalence between the two. Taken together, the two examples illustrate the basic ideas of resolutions and derived functors.
- §4:
- Lie algebras model nonlinearities, infinitesimally. We sketch the central dogma of derived deformation theory, where differential graded (or dg) Lie algebras play, in some sense, the role of resolutions. On the way, we give brief, physics-motivated introductions to the spectrum, the functor of points, and other basic intuitions from algebraic geometry, at a cocktail-party level. The main point is to arrive at the definition of the central object, a local Lie algebra. All technical details are omitted.
- §5:
- A field guide to useful local Lie algebras. We list a bunch of useful examples of local Lie algebras, representing both typical building blocks of field theories and central examples of symmetries. Self-dual fields, spacetime symmetries, and holomorphic symmetries are included, the latter encompassing “higher” Virasoro and Kac–Moody algebras in arbitrary complex dimension.
- §6:
- Current algebras. This is essentially a discussion of the factorization-algebra enhancement of Noether’s theorem proved by Costello and Gwilliam, but it leads to ideas related to Koszul duality, a well-intentioned screed on terminology, some discussion of the operation of gauging, and a bit of ill-advised speculation.
- §7:
- Higher Virasoro algebras in theories. We give an explicit computation of the holomorphic twist of the moduli problem of four-dimensional superconformal structures, obtaining the moduli problem of complex structures on a locally conformally Kähler manifold and, thus, the higher Virasoro algebra. While the result is known [9], the computation is nevertheless illuminating.
1.2. Key Notation and Terminology
- Manifolds are always smooth and oriented; they are usually denoted by M and have dimension d.
- Vector bundles over M are usually denoted by an uppercase Roman letter, such as E.
- Categories (§2.1.3) are denoted by a suggestive name in sans-serif type (e.g., ).
- Groupoids (§2.2.5) are categories in which every map is invertible.
- Sheaves (Section 2.2) are denoted by a calligraphic letter (possibly followed by other letters), and will be used for the sections of a bundle E. An underlined object, as in , denotes the locally constant sheaf with that value.
- (Cochain) complexes are collections of objects indexed (“graded”) by the integers. The grading will be written as a superscript. The total object is defined as the direct sum of its graded components, which exists in any category of objects in which we will consider complexes. Total objects are denoted with a bullet: thus, for example,
- Differential forms are denoted by , and carry the exterior derivative (de Rham differential). They are a sheaf of cochain complexes.
- Densities on a manifold form a sheaf, denoted . The sections of this sheaf can be integrated to produce a number, assuming the integral converges. A density is a top form, twisted by the orientation line bundle; since we assume a choice of orientation for expository purposes, . The Lagrangian is a density.
- Lie algebras will be denoted, as usual, by Fraktur letters like . However, a local Lie algebra—which is a cochain complex of bundles, equipped with a Lie bracket on sections—will be denoted by a sheaf-type letter, usually .
1.3. Acknowledgements
2. A Field Theory as a Sheaf of Spaces
2.1. General Covariance
- 2.1.1. Cartoons of field theories. When one thinks about a semiclassical field theory, the typical package of structures that one has in mind is something like the following:
- 1:
- One first defines a class of spacetime geometries. In essentially every example, these are taken to be smooth manifolds of a fixed dimension d, equipped with all additional geometric data on which the field theory will depend. Among the most important examples are spin structures, (pseudo-)Riemannian metrics, conformal structures, and complex structures.
- 2:
- To each spacetime geometry M, one associates a space of fields , which is often the space of sections of some chosen natural vector bundle. But could also include connections (in gauge theory), maps to some fixed target manifold X (in sigma models), or metrics (in gravity). The space of fields is graded by ; the grading indicates whether the fields are bosons or fermions.
- 3:
- There is a local Lagrangian density, which is a functional , assigning a top form (or, more properly, density) on the spacetime to each possible field configuration in an appropriately local manner. The action functional of the theory is
- 4:
- If the action functional is degenerate, there are nontrivial differential relations (sometimes called “Gauss law constraints” or “Bianchi identities”) between the equations of motion defining , so that we do not get a well-defined boundary value problem. Noether’s second theorem relates these identities to the action of local gauge symmetries on , preserving S. We then pass to the quotient, so that the space we are physically interested in is
- 2.1.2. The reader may object that I am calling such theories “semiclassical”. Every maneuver described above is purely about classical field theory, so why “semi”? The answer has to do with the requirement that the theory is variational. It would have been possible to describe a broader class of field theories, just described by the solution spaces of natural systems of PDEs on the manifold M. But there is no clear way to talk about what the quantization of such a theory should be; that data are encoded, roughly speaking, by the Poisson bracket, whose existence is guaranteed for variational problems. If one wants, one can imagine that generic Newtonian mechanics is classical, but that Lagrangian or Hamiltonian mechanics is semiclassical. We will return to this point later.
- 2.1.3. It is instructive to put this cartoon into a slightly more structured setting. In doing this, I will freely allow myself to use the terminology of categories and functors, but the only deep purpose of this is as a philosophical reminder: One should always ask oneself about the maps!
- 2.1.4. So what maps are appropriate here? The most obvious ones have to do with the fact that fields are locally defined. If is an open subset of some spacetime manifold, and we choose a solution to the equations of motion over M, we can restrict it to N and get a solution to the equations of motion there. So, there should be a restriction map from to and, correspondingly, from to (since we can also restrict an off-shell field configuration). The condition of openness can be thought of as asking that each point has a typical-looking infinitesimal neighborhood; there are no boundaries to deal with.
- 2.1.5. Open embeddings. In fact, these two examples fit together into a single notion. Given some class of smooth, d-dimensional spacetime manifolds equipped with geometric structures as above, a smooth, structure-preserving map is called an open embedding if
- its image is an open subset;
- f is injective. (So f carries N isomorphically onto its image.)
Note that the condition of being structure-preserving and the open embedding condition are independent. It is clear that any open embedding factors into a pair of an isomorphism of N onto , followed by the inclusion . So open embeddings capture both the kinds of maps we identified physically above.
- 2.1.6. Local equivalences. We are working towards identifying an appropriate category that models spacetimes, together with maps that describe local equivalences between spacetimes. (The reader will hopefully recognize this as a more coordinate-free form of the principle of general covariance.) Open embeddings are essentially the right notion, but we will, in fact, work with a slight generalization, motivated by the desire for full locality:
- 2.1.7. Having understood this, we can extend our cartoon from above so as to include both relativistic invariance and locality at the same time:
- 5:
- The assignment
- To any spacetime M, the functor assigns the space of on-shell field configurations on M modulo gauge transformations; the action of the functor on morphisms encodes the restriction of fields and also spacetime symmetries, as explained above. Such a functor is called a presheaf.
- 2.1.8. An example of a field theory. An instructive and relatively uncomplicated example of a field theory in this framework is provided by thinking about a single classical particle moving in the flat n-dimensional space . We emphasize that we are thinking of this theory as a one-dimensional field theory. In general in such formalisms, the degrees of freedom are fields; the parameters labeling points at which those fields can be measured are coordinates on the “spacetime”. In the standard mechanics (whether classical or quantum) of a single particle, the spatial coordinates are physical degrees of freedom that can be measured at any moment in time; the “spacetime”, thus, consists purely of a one-dimensional timeline, whereas the spatial coordinates are promoted to fields.
- 2.1.9. Examples of structures on spacetime. For later use, let us just list some of the most common examples of spacetimes here. The reader is free to ignore any example that feels unfamiliar, at least for the time being.
- 1:
- There is a site of smooth d-dimensional manifolds without any additional structure; its local equivalences are just local diffeomorphisms. A topological field theory is a field theory defined on this site.
- 2:
- If we consider smooth d-manifolds equipped with conformal structures, with local equivalences given by smooth, locally bijective conformal maps, we obtain a site . A conformal field theory is a field theory defined here.
- 3:
- There is a site whose objects are smooth d-manifolds equipped with Riemannian metrics, and whose local equivalences are local isometries. A Euclidean field theory is a field theory defined on this site. Lorentzian field theories are defined analogously, and additional data such as a choice of spin structure can be included in obvious fashion. We emphasize that a Riemannian structure can be profitably thought of as a conformal structure together with a choice of volume form.
- 4:
- We can consider spacetimes that are smooth manifolds of even dimension , equipped with a choice of complex structure. Local equivalences are given by local biholomorphisms. This defines a site of complex n-manifolds; a holomorphic field theory is a field theory defined here.
- Recall that a smooth supermanifold is a smooth d-manifold, equipped with a sheaf of commutative algebras that can locally be identified with a finitely generated exterior algebra (say on k generators) over the smooth functions. A superconformal structure consists, speaking roughly, of a local frame of odd vector fields whose torsion reproduces the structure constants of the supertranslation algebra of flat space. We direct the reader to ([14], chapter 5, §7) for an enlightening early treatment, or to ([9], §2.1.6) for the precise definition we choose to work with. A superconformal structure is the minimal piece of geometric data on which a supersymmetric field theory depends.
- 5:
- There is a site of smooth supermanifolds equipped with a superconformal structure of type . A superconformal field theory is a field theory defined here.
- 6:
- One can also consider sites of smooth supermanifolds equipped with a superconformal structure and additional geometric data. Typically, the additional datum is a chosen section of the Berezinian [15]. A supersymmetric field theory is a field theory defined on such a site.
2.2. Gluing
- 2.2.1. An expository simplification. There are two levels of generality at which one could discuss gluing properties. In the discussion above, we were imagining a field theory as something that assigns a space of field configurations to any appropriately structured spacetime, in a way that is compatible with all local equivalences. As a physicist, this is often the level at which one is implicitly thinking: We know, for example, what it means to place type IIB supergravity on a background of the form , where X is any Calabi–Yau threefold.
- 2.2.2. A category of subsets. To a fixed d-dimensional spacetime M, we can attach a category . The objects of this category are the open subsets of M, and the arrows are the inclusions. Thus, the set of morphisms contains either one element (when ) or zero (otherwise).
- 2.2.3. Gluing conditions. Given two open subsets of spacetime, we can consider the following diagram in :
- 2.2.4. Examples of sheaves. We quickly list just a few basic examples (and classes of examples) of the above constructions here.
- Given an object , the locally constant sheaf takes the value X on any connected open set. The gluing condition can be used to deduce its value on arbitrary open sets; when has direct sums, we find that , with .
- Suppose has a zero object. Then, given a space M and a point , the skyscraper sheaf takes the values
- Given a map of spaces and a sheaf on M, the pushforward is a sheaf on N, defined by taking
- For M a smooth manifold, the smooth sections of any vector bundle over M form a sheaf.
- The space of smooth maps from M into a target space X; to every open set U, we assign , with the obvious notion of restriction. (The fields of the nonlinear sigma model arise in this way.) In fact, this example defines a sheaf on the smooth étale site: To any smooth space, we assign its space of smooth maps into X.
- The solutions to any local partial differential equation define a sheaf.
- A map of vector bundles defines a map between the corresponding sheaves of sections. The kernel and cokernel of such a map are not necessarily vector bundles—but the kernel and cokernel of the map on sections do define sheaves.
- 2.2.5. An important non-example. This example is so important that we will give it special typographical emphasis:
- One easy way of seeing this is the familar Aharonov–Bohm effect [17]. The electromagnetic field is a connection, and its vacuum configurations are described by the condition that its curvature vanishes. Locally, over any open set, any such flat connection is gauge-equivalent to the zero connection. But the space of gauge-equivalence classes of vacua on is of positive dimension, corresponding to the possible holonomies of the gauge field around the excluded region. The corresponding phase effect is observable [18].
- 2.2.6. Moduli spaces. To close this section, we remark briefly on the concept that we have arrived at in the previous example, which is often referred to as a moduli space. At an intuitive level, moduli spaces classify objects: There is one point for each object, and point A is close to point B when object B is, in some appropriate sense, a “small deformation” of object A. This general theme will be ubiquitous in what follows.
3. Resolutions
3.1. From Local to Global
- 3.1.1. Constraints. Thinking about the example of the locally constant sheaf makes it clear that sheaves can contain global information—even though they model locally defined data. Knowing the value of a locally constant function at any point in M determines everything there is to know, if M is connected. There are many examples of this: Knowing a holomorphic function in an open neighborhood of a point determines it in larger open neighborhoods by analytic continuation. Equations of motion in physics are typically well-defined boundary problems, so that knowing the value of a solution on uniquely determines that solution anywhere in U.
- 3.1.2. Gedankenexperimente. Imagine that a space X, for simplicity a d-dimensional smooth manifold, is populated by a collection of local observers. Each of them can observe only some neighborhood of their immediate location, which we assume must be contractible. (This assumption is intuitively reasonable if, for example, we imagine that these regions are defined by some inequality on the geodesic distance. More generally, since any point in a manifold has a neighborhood that locally looks like affine space, we can think of this as a “smallness” assumption on the observers: They should not directly observe any global topology.)
- 3.1.3. What is the result of this procedure at the end of the day? Drawing a diagram of all of the reports of all of the watchmen, we see that we have a diagram of inclusion maps between elements of of the following form:
- 3.1.4. Applying the functor of sections of our sheaf to (12), we obtain a diagram in the target category:
- 3.1.5. Names. We introduce some terminology after the fact. Omitting the first term, the diagram (13)—obtained by applying our sheaf to a diagram showing us how to populate X with local observers, or (equivalently) how to paste X together from contractible open sets—is called the Čech complex, and its cohomology is called the Čech cohomology of the sheaf. When X is a nice enough space, such as a manifold, this cohomology is the same as the sheaf cohomology of , and the higher cohomology groups are the derived functors of the functor of global sections.
3.2. Resolutions by Fine Sheaves
- 3.2.1. Flabbiness. One obvious circumstance that ensures that the sequence (9) is exact is if the restriction map
- 3.2.2. Fineness. The essential issue with continuity is that, given the value of a continuous function at a point, its value at infinitesimally close points (though not at any finite distance) is constrained. This suggests an obvious weakening of the condition of flabbiness: A sheaf on a manifold is called fine if, for any pair of sets , any local section over U, and any closed set , a section over V can be found that agrees with everywhere in K. (We are not being careful about what it means to restrict to a subset that is not open; this is not, strictly speaking, an operation that is a priori meaningful for an arbitrary sheaf. For functions or sections of vector bundles, the notion is intuitively clear; for the general case, we refer to [19,20].)
- 3.2.3. Imposing constraints locally. Our prime example of a sheaf where local information obviously constrains faraway behavior was the locally constant sheaf . Interestingly, this sheaf is a subsheaf of a fine sheaf: There is an obvious inclusion map
- 3.2.4. A smooth function is locally constant precisely when each of its derivatives vanishes at every point. Succinctly, f is locally constant when it is annihilated by the exterior derivative operator d. This tells us the first step of our resolution: We can take
- 3.2.5. Local constraints replace local observers. We now have an easy way to see the fact that the Čech cohomology groups we associated to the sheaf on a space X above are nothing other than the de Rham cohomology groups of X. (This pattern of argument is extremely common, and goes back in this example to [22].)
3.3. Resolutions of Invariants
- 3.3.1. Let be a Lie algebra. For simplicity, we work in the category of linear -representations over a field of characteristic zero (say or ).
- 3.3.2. The fundamental observation in the case of sheaves had to do with studying the basic pasting diagram (7)—an analogue, in spaces, of a short exact sequence. Applying the functor defined by the sheaf, we observed that the resulting sequence (9) was necessarily exact on the left, but not on the right. The failure of (9) to be exact was the origin of the higher cohomology of the sheaf.
- 3.3.3. Derived invariants. By now, the path should be clear. Given a -module A, we should try and impose the condition of being -invariant explicitly, “at the cochain level”, rather than strictly. By doing this, we should arrive at a recipe for “pasting together” the subspace out of copies of A.
- 3.3.4. It should also be apparent to the reader who has seen the BRST formalism for gauge theories [24,25] that we are recovering the basic ingredient of that formalism in a much more general setting. If we take the module A to be the smooth functions on some space, we can think of the Lie algebra cochains as being functions on a “graded manifold” . In cohomological degree zero, we recover -invariant functions, which are a model for the functions on the quotient space of X by the infinitesimal action of . In the context of the BRST formalism, the new degrees of freedom responsible for taking derived -invariants, which consist of the symmetry generators placed in degree , are the fields usually called “ghosts.”
- 3.3.5. A word on gradings. In the context of physics, every system is equipped with a grading by , called fermion parity or intrinsic parity. Physical degrees of freedom can be either bosons, represented by commuting fields, or fermions, represented by anticommuting fields.
- 3.3.6. It turns out that thinking about Lie algebra cochains will allow us not just to construct a derived model for gauge invariants, but a derived model for an entire perturbative field theory. A version of this construction provides one possible answer to the open question from above (§2.1.7) about which notion of spaces we want to work with. We turn to this now, beginning with a bit of a detour through intuitive basic algebraic geometry.
4. Local Lie Algebras and Formal Moduli Problems
4.1. Intuitions from Algebraic Geometry
- 4.1.1. A basic observation. We start at the very beginning:
- 4.1.2. Measurements. Imagine that we are interested in a physical system with algebra of observables A. To perform a simple measurement, we select an element , and inform an experimentalist that we are interested in knowing what value f takes. After taking appropriate action in the world, the experimentalist returns some numerical value a. (Better yet, we could engage with the world of phenomena ourselves. For the sake of brevity, though, let me refrain from turning down that road.)
- 4.1.3. Numerical measurements. (The author owes many of the following explanations to discussions with Minhyong Kim, whom he acknowledges with gratitude.) One important class of ideals correspond, intuitively, to “measurements valued in numbers.” Since there are no obstructions to simultaneous measurement in classical mechanics, such measurements should specify a numerical value for all observables. We expect that numerical measurements “determine the state”, so we would like to arrive at a notion that captures the idea of a “point” in the state space.
- 4.1.4. The generic point. There is an important distinction to make at this juncture, which is most easily seen by considering the simple example of polynomials in one variable, . We think of these as polynomial functions on . What are the prime ideals here?
- 4.1.5. Maximal measurements. Having understood this, we have arrived at the correct intuition. A point in consists of a possible numerical measurement, which may or may not be maximal. (It may provide incomplete numerical information.) If A comes to us as the algebra of functions on some space X, we can imagine each point as some subspace (of whatever dimension) of X that can be cut out by numerical measurements. In particular, the entire space X (the “generic point”) is a point of .
- 4.1.6. Unreliable experimentalists. Returning to Idea 1, it is natural to ask about the physical relevance or intuitive meaning of the other ideals that we have been ignoring. Again, we will just think about the basic example .
4.2. The Functor of Points
- 4.2.1. The definition. Having absorbed our notion of “point” from above, there are a few powerful generalizations of the basic idea that we can consider. The essential idea, already sketched above, is that we can imagine measurements of different types, according to where the “evaluation map” is valued. For instance, we can think of as being the set of real points of A, as being the set of complex points, and so on. (If A is an algebra over R, such maps will automatically be surjective and, thus, define maximal ideals.) Since a field k has no nontrivial ideals, consists of only one point, and maps from A to k can be thought of dually as maps of spaces of the form
- 4.2.2. Abstracting away from this, it is useful to think about such a functor of points as being a generalization of a space of the form . Certainly, any space that is at least locally modeled by some defines a functor of points. But not any functor deserves to be called a functor of points: Thinking of the example of the sigma model in §2.2.4, it is clear that we should ask for some sort of gluing condition on the set of test spaces, giving us a notion of compatibility between the space of Z-points and the spaces of -points when the test spaces form a “covering” of Z in an appropriate sense. Interesting generalizations of the notion of space arise when a functor has all of the properties of a functor of points—in particular, satisfies locality, in the form of a “gluing” or “descent” condition—but is not represented by any space.
- 4.2.3. At this point, one might now attempt to describe the “space” of fields in a field theory on a manifold M using a sheaf of groupoids on some category of test spaces. To a test space Z, we assign the groupoid of families of gauge field configurations on M parameterized by Z.
- 4.2.4. The second step involves pinning down the relevant category of test objects, which ends up being the following:
- 4.2.5. Formal moduli problems. At this point, the reader should hopefully be able to guess the outline of the definition of a formal moduli problem, which we quote here for the sake of completeness. Let denote the category of dg local Artin algebras, as defined above, and the category of simplicial sets ([7], Definition 4.9).
4.3. Deformation Functors
- 4.3.1. We now want to lay the final stone in the winding path of intuitions that is bringing us to the notion of a local Lie algebra. The essential idea involved is both old and deep. I quote it here in Lurie’s formulation:
- 4.3.2. Applying this general philosophy in the context of field theory, Costello and Gwilliam give a definition of a local Lie algebra which is adapted to describe any natural formal moduli problem defined by local partial differential equations on a manifold.
4.4. Variational Formal Moduli Problems
- 4.4.1. We have now completed motivating the definition of a local Lie algebra, and built up intuition as to why the definition is correct, why it is powerful, and what it can be applied to. It remains to understand which formal moduli problems correspond to perturbative Lagrangian field theories. In physics, the development of a derived approach to Lagrangian field theories was pioneered in groundbreaking work by Batalin and Vilkovisky [44,45,46], and still goes by the name of the Batalin–Vilkovisky formalism. We deeply regret not being able to discuss it at greater length here.
5. Examples
5.1. Introductory Remarks
- 5.1.1. In what follows, I will try to use names for local Lie algebras that remind the reader of the formal moduli problems they represent. Mimicking the typical notation for Lie groups and Lie algebras, I will denote formal moduli problems with lowercase letters: for example, “” for infinitesimal deformations of a fixed flat G-connection. The corresponding full moduli problem, consisting of the space of all flat G-connections, would be denoted . We emphasize again that a formal moduli problem involves a choice of basepoint, analogous to the choice of vacuum in perturbative quantum field theory; our notation often does not indicate this choice explicitly, but it is always present.
- 5.1.2. Anomalies. The cohomological approach to anomalies, which identifies them as elements in a particular BRST cohomology group associated to the local Lie algebra representing the symmetry, has a long history. Implicitly, it goes back as far as the celebrated Wess–Zumino consistency conditions [47]; the subject was developed rapidly by the Italian school in the 1908s ([48], among others), and is by now very much a part of the standard lore. We cannot hope to review it here, and refer the reader to any of the comprehensive treatments in the literature [49,50,51].
- 5.1.3. Sphere algebras. It is worth emphasizing again that local Lie algebras are, in particular, sheaves of Lie algebras on an entire class of manifolds. They, therefore, contain much more information than a single symmetry algebra, and should not be confused with the values they take on flat space, or on any particular geometry.
5.2. Locally Constant Symmetries
- 5.2.1. Let be a finite-dimensional real or complex Lie algebra, and G a Lie group with . Let denote the site of smooth d-manifolds equipped with principal G-bundles P carrying a chosen connection. The connection induces a derivation ∇ of degree in the complex of de Rham forms valued in any vector bundle associated to P, in particular, in the adjoint bundle , which is a bundle of Lie algebras. The curvature is the tensor ; when the connection is flat, ∇ is a square-zero differential.
- 5.2.2. We now work again over the site .
- 5.2.3. Our third formal moduli problem is, in a sense, intermediate between and . For the sake of brevity, we describe it only near the trivial connection in the trivial principle G-bundle on M.
5.3. Higher-Form Symmetries
- 5.3.1. Connections in higher abelian gerbes. We give some relatively obvious generalizations of the abelian versions of the above formal moduli problems to the setting of higher-form symmetry.
5.4. Spacetime Symmetries
- 5.4.1. Riemannian structures. The local dg Lie algebra met controlling deformations of (equivalence classes of) Riemannian metrics is defined on any smooth manifold M with Riemannian metric g. It takes the form
- 5.4.2. Riemannian structures on the line. We write a degenerate example of this formal moduli problem (parallel to §2.1.8) explicitly in coordinates: the one-dimensional case. Working with respect to a generic coordinate s on the generic line, a metric is given by , where f is a globally nonvanishing function and denotes the tensor square (automatically symmetric in dimension one). As discussed above, we can choose an adapted coordinate t such that . A vector field is given by an expression of the form , with x again a function, and a perturbation of the metric by a tensor . The differential sends to the element ; the nonvanishing Lie brackets are
- 5.4.3. Conformal structures. Working on the same site, smooth manifolds equipped with Riemannian metrics, we can define a local dg Lie algebra conf controlling deformations of conformal equivalence classes of metrics. It takes the form
- 5.4.4. Frames and G-structures. There is a model for the formal moduli problem of G-frames on d-manifolds, where G is a Lie group equipped with a map to . In the case where we choose or , the resulting moduli problem is equivalent to the moduli problem of metrics or of conformal structures, respectively. We refer to [9] for details.
- 5.4.5. Supersymmetries. We make the discussion of superconformal structures, briefly mentioned in §2.1.9, a bit more precise. Such a structure consists of a subbundle D of the tangent bundle of a smooth supermanifold M, of maximal odd dimension, such that the symbol, defined on the associated graded bundle by taking the image of the Lie bracket of vector fields in D under the projection to , is a locally trivial bundle of super Lie algebras. The typical fiber of this bundle, which is a two-step nilpotent super Lie algebra of the same dimension as M, is called the type of the superconformal structure. In physical terms, is the supertranslation algebra, and D specifies the vector fields on superspace that allow one to think of it as locally modeled by the flat superspace .
5.5. Holomorphic Symmetries
- 5.5.1. The local (higher) Kac–Moody algebra. The constructions we explore here go back to [61,62] and were related to twists of current multiplets in supersymmetric theories in [63]. We work on a complex manifold X. Recall that a holomorphic vector bundle on X consists of a smooth vector bundle E over , equipped with a Dolbeault operator
- Leibniz: For any section of E and any smooth function f, we have
- Cauchy–Riemann: After extending to as a derivation, .
- 5.5.2. The local (higher) Virasoro algebra. We work on the site of complex n-manifolds.
5.6. Holomorphic Factorization
- 5.6.1. Riemann surfaces. Let X be a smooth, oriented two-dimensional manifold, and equip it with a Riemannian metric g. We can think of g as a pair, consisting of a conformal structure (denoted ) and a volume form (denoted ). Recall that these data determine a Hodge star operator,
- 5.6.2. Perturbative equivalence. Since we have already seen that conformal structures and complex structures are the same, we must find an equivalence between the corresponding perturbative formal moduli problems as defined in the previous sections. Let us see how this works out explicitly. For simplicity of exposition, we work in the neighborhood of the flat metric on flat space. (As we did in §5.4.3, we think of both moduli problems as being defined on .)
- 5.6.3. We exhibit a map of dg Lie algebras from to :
- A holomorphic vector field X is sent via the identity operator to the summand and via the negative of the divergence operator to the summand in .
- A Beltrami differential is sent to the antiholomorphic quadratic differential .
- 5.6.4. Weyl and Virasoro cocycles. We now pull back the explicit cocycle
6. Observables and Currents
6.1. Motivating Example: Hamiltonian Mechanics
- 6.1.1. Infinitesimal symmetries of phase space. Let be a symplectic manifold. Recall that the space of symplectic vector fields on M is given by
- 6.1.2. A word on quantization. For orientation, we offer a brief aside on quantization, which will probably be apparent to the reader anyway. One should think of a Poisson algebra as the “semiclassical residue” of a one-parameter family of associative algebras, parameterized by ℏ. The Poisson algebra we have constructed consists of (polynomial) functions on the linear space , equipped with the natural (Kirillov–Kostant–Souriau) Poisson bracket. Its quantization is just the universal enveloping algebra ; in quantum mechanics, we would have asked for a projective representation of on Hilbert space, which amounts to a map from to the algebra of linear operators. (Unitarity is, of course, important, but we will not discuss it here.)
- 6.1.3. Duality. There is an important sense in which the algebra of -currents and the algebra of -ghosts are dual to one another. The relationship is known as Koszul duality, and is witnessed by the fact that admits a differential that makes it quasi-isomorphic to the ground field. Morally speaking, the two factorization algebras we will soon associate to a local Lie algebra are also an instance of a Koszul dual pair. We do not have anything more specific to say about this here, but it is worth mentioning.
6.2. The Factorization Noether Theorem
- 6.2.1. We first give a schematic explanation to justify the shifts in cohomological degree appearing in the statement, which may be unfamiliar. Recall that the moduli space of fields in a BV theory is equipped with a -shifted symplectic structure . Correspondingly, modulo various subtleties, the observables should be equipped with a -shifted Poisson bracket. One expects a surjective map
- 6.2.2. As we have seen above, a local Lie algebra defines a formal moduli problem, and a variational local Lie algebra defines a perturbative Lagrangian field theory. The observables are given by the factorization algebra
- 6.2.3. Noether’s theorem via descent. It deserves a bit of emphasis that the standard proof of Noether’s theorem already implicitly appeals to the structure of a derived replacement of the sheaf , namely, by the local Lie algebra controlling deformations of the trivial connection. In fact, even careful versions of the standard proofs appeal to a version of homological descent, but in the variational bicomplex rather than in a local resolution; the reader interested in learning more is referred to the careful discussion in ([70], §§2.6–8). We are not aware of any comparably careful treatment incorporating both spacetime descent and BV descent explicitly.
- 6.2.4. Visualizing the structure of . We now turn to seeing how this intuition is witnessed at the level of the factorization algebra of currents. Given a local Lie algebra, the generators of its currents are . It is useful for the intuition to recall that there is a pairing between compactly supported sections of the bundle L and distributional sections of . Here,
- 6.2.5. Examples: global symmetries. Recall the three formal moduli problems extending the constant sheaf that we constructed above in Section 5.2. We visualize the structure of the currents for each as performed in the previous section. The currents for take the form
- 6.2.6. Examples: the stress tensor. The currents for the moduli problem of metrics take the form
6.3. Disambiguation
- 6.3.1. Local symmetries versus gauge invariances. There is sometimes an unfortunate habit of conflating the term “local symmetry” with the term “gauge symmetry”. The term “local” is drastically overburdened in any case. For the purposes of disambiguation, we sketch the typical speech patterns, which (in our experience) are roughly as follows:
- Every symmetry of a field theory, or, at least, every infinitesimal symmetry, is tacitly assumed to act in a local manner. (Thus, it is encoded by an action of a sheaf of Lie algebras.) Whether or not a symmetry is called “local” has nothing to do with this.
- A symmetry by a locally constant sheaf of Lie algebras—in other words, a symmetry parameterized by functions that are constrained to be locally constant—is called a “global symmetry.”
- A symmetry by a locally free sheaf of Lie algebras—in other words, a symmetry parameterized by one or more unconstrained smooth functions—is called a “local symmetry.”
- A “gauge symmetry” is used to refer to a group action which, as in our cartoon in Section 2, is actually part of a presentation of the space of field configurations as a quotient of some other space.
- In standard situations, Noether’s second theorem [71] shows that symmetries of the action by a locally free sheaf of Lie algebras correspond one-to-one to nontrivial differential relations between the resulting variational equations of motion. When such relations are present, the variational principle does not define a well-defined boundary value problem. The typical remedy is to pass to studying the quotient space of the solutions by the corresponding local symmetry, which is, thus, invariably treated as a gauge invariance. This is the root cause of the tendency to conflate the two terms.
- Working with local Lie algebras offers a lot of clarity on many of these issues, but leads to new terminological complexities. In this article, I do my best to use language in a way that is, at least, largely internally consistent. Here are some principles of usage I hope I have managed to adhere to:
- 1:
- I will exclusively use the term “gauge symmetry” for the action of a group (or Lie algebra) appearing in a description of the sheaf of physical fields as a quotient. This is consistent with standard usage. The gauge symmetries of a theory are not a subset of its symmetries. Where possible, I will, thus, try and refer to “gauge invariances” rather than “gauge symmetries”, though I cannot promise consistency.
- 2:
- “Local” and “gauge” are distinct terms and are not to be equated. In particular, “local Lie algebra” is a general term with a fixed meaning (Definition 4) and does not bear any relation to the distinctions outlined above.
- 6.3.2. The dogma of the ghost in the machine. It is common in expositions of the BRST formalism to emphasize that the ghost fields that one introduces are “a trick”, or perhaps (in some vague sense of the word) “formal”—in any case, certainly “unphysical”. While the narrow point that is being made is not wrong, the would-be ontology that lies behind it is misleading and breeds confusion down the road. (The reader who has time to spare might investigate ([73], §1.2).)
- 6.3.3. Gauging. Recall that actions of a Lie algebra on another Lie algebra correspond to split extensions of the form
- 1:
- Formulate the symmetry in terms of a local Lie algebra . When the symmetry is by a locally constant sheaf of Lie algebras , is typically the moduli problem of G-connections.
- 2:
- Identify a -shifted symplectic formal moduli problem that maps to .
- 3:
- The semidirect product is now itself a variational formal moduli problem, representing the total space of the fibration. After gauging, we study this formal moduli problem, regarding it as a theory in and of itself.
- 6.3.4. Completeness (no global symmetries). In the context of recent attempts to axiomatize the properties of the set of low-energy effective field theories that can arise under the assumption of a consistent coupling to gravity, there has been much discussion of the conjectured absence of global symmetries. In the context of string theory, this absence is well established, originating as a “folk theorem” known, at least, to ’t Hooft, Susskind, and Witten, and seemingly first discussed in the literature in [74]. General arguments connecting black holes to global symmetry violation go back even further [75].
Mathematics was not sufficiently refined in 1917 to cleave apart the demands for “no prior geometry” and for a geometric, coordinate-independent formulation of physics. Einstein described both demands by a single phrase, “general covariance”. The “no prior geometry” demand actually fathered general relativity, but by doing so anonymously, disguised as “general covariance”, it also fathered half a century of confusion.
- 1:
- Carefully formulate a well-defined notion of “completeness” for local Lie algebras, and understand its meaning for the corresponding formal moduli problems. (They should, in some sense, not occur as the fiber of any nontrivial family over another formal moduli problem on the site of manifolds.)
- 2:
- Argue that any complete local Lie algebra does, in fact, describe an formal moduli problem. Alternatively, identify a set of sufficient conditions for this to be the case, or identify an obstruction to the existence of an structure.
7. Higher Virasoro Algebras in Supersymmetric Theories
7.1. Minimal Conformal Supergravity in Four Dimensions
- 7.1.1. Let M be a smooth oriented four-manifold, and let denote its complexified tangent bundle. We fix two complex vector bundles on M, denoted , together with an isomorphism
- 7.1.2. Denote the Chern roots of the bundles by and , respectively. The Chern roots of the bundle are, then,
- 7.1.3. A choice of superconformal structure of type determines a reduction in the structure group to ; the corresponding principal bundle is the bundle of adapted frames. (For early work on G-structures on supermanifolds and adapted frames, see [79,80].) For structures in four dimensions, the corresponding reduction in the structure group is along the map
- 7.1.4. To perform the twist, we first put a -graded structure on our supermanifold, placing in cohomological degree and in cohomological degree . The theory admits a holomorphic twist precisely when one of the two bundles—say , to conform to our choice of regrading—admits a globally nonvanishing section. This means that splits as a sum of lines
- 7.1.5. A choice of splitting of as a sum of lines implies a further reduction of the structure group, and fixing a section of the trivial sub-line (a twisting supercharge) reduces down to
- 7.1.6. Following [9], the underlying vector bundle of the moduli problem of four-dimensional superconformal structures takes the form
- 7.1.7. We observe that the generators that survive the twist correspond precisely to the generators of the Dolbeault resolution of holomorphic vector fields, which is given by the dg Lie algebra . Generators in degree zero arise from the holomorphic part of the smooth vector fields that implement the quotient by diffeomorphisms in the original moduli problem (112).
7.2. Anomalies and Cocycles
- 7.2.1. In four dimensions, there are two independent anomaly cocycles in . The a cocycle is defined by the product of the ghost for Weyl rescaling with the Euler density, evaluated on the deformed metric . In this dimension, the Euler density is quadratic in the Ricci curvature. For the c cocycle, the Euler density is replaced by the square of the conformally invariant Weyl tensor of .
Funding
Data Availability Statement
Conflicts of Interest
References
- Costello, K.; Gwilliam, O. Factorization Algebras in Quantum Field Theory II; Number 41 in New Mathematical Monographs; Cambridge University Press: Cambridge, UK, 2021. [Google Scholar]
- Pantev, T.; Toën, B.; Vaquié, M.; Vezzosi, G. Shifted symplectic structures. Publ. Mathématiques de l’IHÉS 2013, 117, 271–328. [Google Scholar] [CrossRef]
- Vakil, R. The Rising Sea; Princeton University Press: Princeton, NJ, USA, 2024. [Google Scholar]
- Hamilton, W.R. On a General Method of Expressing the Paths of Light, & of the Planets, by the Coefficients of a Characteristic Function; P. Dixon Hardy: Dublin, Ireland, 1833. [Google Scholar]
- Costello, K.; Gwilliam, O. Factorization Algebras in Quantum Field Theory I; Number 31 in New Mathematical Monographs; Cambridge University Press: Cambridge, UK, 2017. [Google Scholar]
- Freed, D.S. Lectures on topological quantum field theory. In Proceedings of the GIFT Summer School on Recent Problems in Mathematical Physics, Salamanca, Spain, 15–27 June 1992. [Google Scholar]
- Freed, D.S.; Hopkins, M. Chern–Weil forms and abstract homotopy theory. Bull. Am. Math. Soc. 2013, 50, 431–468. [Google Scholar] [CrossRef]
- Freed, D.S. What is an anomaly? arXiv 2023, arXiv:2307.08147. [Google Scholar]
- Hahner, F.; Raghavendran, S.; Saberi, I.; Williams, B.R. Local superconformal algebras. arXiv 2024, arXiv:2410.08176. [Google Scholar] [CrossRef]
- Mac Lane, S. Categories for the Working Mathematician, 2nd ed.; Number 5 in Graduate Texts in Mathematics; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar]
- Coecke, B.; Paquette, E.O. Categories for the practicing physicist. In New Structures for Physics; Coecke, B., Ed.; Number 813 in Lecture Notes in Physics; Springer: Berlin/Heidelberg, Germany, 2010; pp. 173–286. [Google Scholar]
- Behrend, K.; Xu, P. Differentiable stacks and gerbes. J. Symplectic Geom. 2011, 9, 285–341. [Google Scholar] [CrossRef]
- Milne, J.S. Lectures on Étale Cohomology; Number 33 in Princeton Mathematical Series; Princeton University Press: Princeton, NJ, USA, 1980. [Google Scholar]
- Manin, Y.I. Gauge Field Theory and Complex Geometry; Number 289 in Grundlehren der mathematischen Wissenschaften; Springer: Berlin/Heidelberg, Germany, 1984. [Google Scholar]
- Deligne, P. Superspace approaches to supergravity. In Quantum Fields and Strings: A Course for Mathematicians; Deligne, P., Etingof, P., Freed, D.S., Jeffrey, L.C., Kazhdan, D., Morgan, J.W., Morrison, D.R., Witten, E., Eds.; American Mathematical Society: Providence, RI, USA, 1999; Volume 2. [Google Scholar]
- Giraud, J. Analysis Situs; Number 256 in Séminaire Bourbaki 1962/63; Secrétariat Mathématique, Université Paris: Paris, France, 1964. [Google Scholar]
- Aharonov, Y.; Bohm, D. Significance of electromagnetic potentials in the quantum theory. Phys. Rev. 1959, 115, 485. [Google Scholar] [CrossRef]
- Tonomura, A.; Osakabe, N.; Matsuda, T.; Kawasaki, T.; Endo, J.; Yano, S.; Yamada, H. Evidence for Aharonov–Bohm effect with magnetic field completely shielded from electron wave. Phys. Rev. Lett. 1986, 56, 792–795. [Google Scholar] [CrossRef] [PubMed]
- Godement, R. Topologie Algébrique et Théorie des Faisceaux; Number 13 in Publications de l’Institut de Mathématique de l’Université de Strasbourg; Hermann: Paris, France, 1958. [Google Scholar]
- Bredon, G.E. Sheaf theory. In Graduate Texts in Mathematics; Springer: Berlin/Heidelberg, Germany, 2012; Volume 170. [Google Scholar]
- Swan, R.G. Vector bundles and projective modules. Trans. Am. Math. Soc. 1962, 105, 264–277. [Google Scholar] [CrossRef]
- Weil, A. Sur les théorèmes de de Rham. Comment. Math. Helv. 1952, 26, 119–145. [Google Scholar] [CrossRef]
- Chevalley, C.; Eilenberg, S. Cohomology theory of Lie groups and Lie algebras. Trans. Am. Math. Soc. 1948, 63, 85–124. [Google Scholar] [CrossRef]
- Becchi, C.; Rouet, A.; Stora, R. Renormalization of the abelian Higgs–Kibble model. Commun. Math. Phys. 1975, 42, 127–162. [Google Scholar] [CrossRef]
- Tyutin, I.V. Gauge invariance in field theory and statistical physics in operator formalism. Lebedev Phys. Inst. 1975, 39. [Google Scholar]
- Deligne, P.; Morgan, J.W. Notes on supersymmetry (following Joseph Bernstein). In Quantum Fields and Strings: A Course for Mathematicians; Deligne, P., Etingof, P., Freed, D.S., Jeffrey, L.C., Kazhdan, D., Morgan, J.W., Morrison, D.R., Witten, E., Eds.; American Mathematical Society: Providence, RI, USA, 1999; Volume 1, pp. 41–98. [Google Scholar]
- Raghavendran, S.; Saberi, I.; Williams, B.R. Twisted eleven-dimensional supergravity. Commun. Math. Phys. 2023, 402, 1103–1166. [Google Scholar] [CrossRef]
- Manetti, M. Differential graded Lie algebras and formal deformation theory. In Algebraic Geometry: Seattle 2005; Number 80 in Proceedings of Symposia in Pure Mathematics; AMS: Amsterdam, The Netherlands, 2009; pp. 785–810. [Google Scholar]
- Atiyah, M.F.; MacDonald, I.G. Introduction to Commutative Algebra; Westview Press: Boulder, CO, USA, 1969. [Google Scholar]
- Vezzosi, G. What is …a derived stack? Not. Am. Math. Soc. 2011, 58, 955–958. [Google Scholar]
- The Stacks Project: An Open Source Textbook and Reference Work on Algebraic Geometry. Available online: https://stacks.math.columbia.edu (accessed on 4 July 2025).
- Lurie, J. Derived Algebraic Geometry X: Formal Moduli Problems. Available online: https://www.math.ias.edu/~lurie/papers/DAG-X.pdf (accessed on 4 July 2025).
- Quillen, D. Rational homotopy theory. Ann. Math. 1969, 90, 205–295. [Google Scholar] [CrossRef]
- Sullivan, D. Infinitesimal computations in topology. Publ. Mathématiques de l’IHÉS 1977, 47, 269–331. [Google Scholar] [CrossRef]
- Kodaira, K.; Spencer, D.C. On deformations of complex analytic structures, I–II. Ann. Math. 1958, 67, 328–466. [Google Scholar] [CrossRef]
- Nijenhuis, A.; Richardson, R.W., Jr. Cohomology and deformations in graded Lie algebras. Bull. Am. Math. Soc. 1966, 72, 1–29. [Google Scholar] [CrossRef]
- Deligne, P. Letter to John Millson. 1987. Available online: https://publications.ias.edu/sites/default/files/millson.pdf (accessed on 4 July 2025).
- Drinfeld, V. A letter from Kharkov to Moscow (1988). Ems Surv. Math. Sci. 2014, 1, 241–248. [Google Scholar] [CrossRef]
- Pridham, J.P. Unifying derived deformation theories. Adv. Math. 2010, 224, 772–826. [Google Scholar] [CrossRef]
- Hinich, V. DG coalgebras as formal stacks. J. Pure Appl. Algebra 2001, 162, 209–250. [Google Scholar] [CrossRef]
- Getzler, E. Lie theory for nilpotent L∞ algebras. Ann. Math. 2009, 170, 271–301. [Google Scholar] [CrossRef]
- Calaque, D.; Grivaux, J. Formal moduli problems and formal derived stacks. Panoramas Synthèses 2021, 55, 85–145. [Google Scholar]
- Vallette, B. Algebra + homotopy = operad. Symplectic Poisson. Noncommut. Geom. 2014, 62, 229–290. [Google Scholar]
- Batalin, I.A.; Vilkovisky, G.A. Gauge algebra and quantization. Phys. Lett. B 1981, 102, 27–31. [Google Scholar] [CrossRef]
- Batalin, I.A.; Vilkovisky, G.A. Closure of the gauge algebra, generalized Lie equations, and Feynman rules. Nucl. Phys. B 1984, 234, 106–124. [Google Scholar] [CrossRef]
- Batalin, I.A.; Vilkovisky, G.A. Existence theorem for gauge algebra. J. Math. Phys. 1985, 26, 172–184. [Google Scholar] [CrossRef]
- Wess, J.; Zumino, B. Consequences of anomalous Ward identities. Phys. Lett. B 1971, 37, 95–97. [Google Scholar] [CrossRef]
- Bonora, L.; Cotta-Ramusino, P. Some remarks on BRS transformations, anomalies, and the cohomology of the Lie algebra of the group of gauge transformations. Commun. Math. Phys. 1983, 87, 589–603. [Google Scholar] [CrossRef]
- Barnich, G.; Brandt, F.; Henneaux, M. Local BRST cohomology in gauge theories. Phys. Rep. 2000, 338, 439–569. [Google Scholar] [CrossRef]
- Bertlmann, R.A. Anomalies in Quantum Field Theory; Number 91 in International Series of Monographs on Physics; Oxford University Press: Oxford, UK, 2000. [Google Scholar]
- Bonora, L. Fermions and Anomalies in Quantum Field Theories; Theoretical and Mathematical Physics; Springer: Berlin/Heidelberg, Germany, 2023. [Google Scholar]
- Adler, S.L. Axial-vector vertex in spinor electrodynamics. Phys. Rev. 1969, 177, 2426. [Google Scholar] [CrossRef]
- Bell, J.S.; Jackiw, R.W. A PCAC puzzle: π0→ γγ in the σ-model. Il Nuovo C. A 1969, 60, 47–61. [Google Scholar] [CrossRef]
- Brylinski, J.L. Loop Spaces, Characteristic Classes, and Geometric Quantization; Number 107 in Progress in Mathematics; Springer: Berlin/Heidelberg, Germany, 2009. [Google Scholar]
- Saberi, I.; Williams, B.R. Constraints in the BV formalism: Six-dimensional supersymmetry and its twists. Adv. Math. 2023, 412, 108806. [Google Scholar] [CrossRef]
- Butson, D.; Yoo, P. Degenerate classical field theories and boundary theories. arXiv 2016, arXiv:1611.00311. [Google Scholar]
- Bonora, L.; Cotta-Ramusino, P.; Reina, C. Conformal anomaly and cohomology. Phys. Lett. B 1983, 126, 305–308. [Google Scholar] [CrossRef]
- Bonora, L.; Pasti, P.; Bregola, M. Weyl cocycles. Class. Quantum Gravity 1986, 3, 635. [Google Scholar] [CrossRef]
- Deser, S.; Schwimmer, A. Geometric classification of conformal anomalies in arbitrary dimensions. Phys. Lett. B 1993, 309, 279–284. [Google Scholar] [CrossRef]
- Boulanger, N. General solutions of the Wess–Zumino consistency condition for the Weyl anomalies. J. High Energy Phys. 2007, 2007, 069. [Google Scholar] [CrossRef]
- Gwilliam, O.; Williams, B.R. Higher Kac–Moody algebras and symmetries of holomorphic field theories. Adv. Theor. Math. Phys. 2021, 25, 129–239. [Google Scholar] [CrossRef]
- Faonte, G.; Hennion, B.; Kapranov, M. Higher Kac–Moody algebras and moduli spaces of G-bundles. Adv. Math. 2019, 346, 389–466. [Google Scholar] [CrossRef]
- Saberi, I.; Williams, B.R. Twisted characters and holomorphic symmetries. Lett. Math. Phys. 2020, 110, 2779–2853. [Google Scholar] [CrossRef]
- Williams, B.R. A complex geometric perspective on the a and c anomalies. arXiv 2024, arXiv:2404.16128. [Google Scholar] [CrossRef]
- Williams, B.R. The local cohomology of vector fields. arXiv 2024, arXiv:2405.05174. [Google Scholar] [CrossRef]
- Gel’fand, I.M.; Fuks, D.B. Cohomology of the Lie algebra of formal vector fields. Math. USSR–Izvestiya 1970, 4, 327. [Google Scholar] [CrossRef]
- Bott, R. On characteristic classes in the framework of Gelfand–Fuks cohomology. Astérisque 1976, 32–33, 113–139. [Google Scholar]
- Gel’fand, I.M.; Fuks, D.B. The cohomologies of the Lie algebra of the vector fields in a circle. Funct. Anal. Appl. 1968, 2, 342–343. [Google Scholar] [CrossRef]
- Knudsen, B. Higher enveloping algebras. Geom. Topol. 2018, 22, 4013–4066. [Google Scholar] [CrossRef]
- Deligne, P.; Freed, D.S. Classical field theory. In Quantum Fields and Strings: A Course for Mathematicians; Deligne, P., Etingof, P., Freed, D.S., Jeffrey, L.C., Kazhdan, D., Morgan, J.W., Morrison, D.R., Witten, E., Eds.; American Mathematical Society: Providence, RI, USA, 1999; Volume 1, pp. 137–226. [Google Scholar]
- Noether, E. Invariante Variationsprobleme. Nachrichten Ges. Wiss. Göttingen Math.-Phys. Kl. 1918, 1918, 235–257. [Google Scholar]
- Brunner, I.; Carqueville, N.; Plencner, D. Orbifolds and topological defects. Commun. Math. Phys. 2014, 332, 669–712. [Google Scholar] [CrossRef]
- Ryle, G. The Concept of Mind; Hutchinson: London, UK, 1949. [Google Scholar]
- Banks, T.; Dixon, L.J. Constraints on string vacua with spacetime supersymmetry. Nucl. Phys. B 1988, 307, 93–108. [Google Scholar] [CrossRef]
- Zeldovich, Y.B. A new type of radioactive decay: Gravitational annihilation of baryons. Phys. Lett. A 1976, 59, 254. [Google Scholar] [CrossRef]
- Misner, C.W.; Thorne, K.S.; Wheeler, J.A. Gravitation; Macmillan: New York, NY, USA, 1973. [Google Scholar]
- Witten, E. Quantum field theory and the Jones polynomial. Commun. Math. Phys. 1989, 121, 351–399. [Google Scholar] [CrossRef]
- Costello, K. Topological conformal field theories and Calabi–Yau categories. Adv. Math. 2007, 210, 165–214. [Google Scholar] [CrossRef]
- Schwarz, A.S. Supergravity, complex geometry, and G-structures. Commun. Math. Phys. 1982, 87, 37–63. [Google Scholar] [CrossRef]
- Fujio, M. Super G-structures of finite type. Osaka J. Math. 1991, 28, 163–211. [Google Scholar]
degree −1 | |
degree 0 | |
degree 1 | |
degree 2 |
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Saberi, I. Notes on Derived Deformation Theory for Field Theories and Their Symmetries. Symmetry 2025, 17, 1172. https://doi.org/10.3390/sym17081172
Saberi I. Notes on Derived Deformation Theory for Field Theories and Their Symmetries. Symmetry. 2025; 17(8):1172. https://doi.org/10.3390/sym17081172
Chicago/Turabian StyleSaberi, Ingmar. 2025. "Notes on Derived Deformation Theory for Field Theories and Their Symmetries" Symmetry 17, no. 8: 1172. https://doi.org/10.3390/sym17081172
APA StyleSaberi, I. (2025). Notes on Derived Deformation Theory for Field Theories and Their Symmetries. Symmetry, 17(8), 1172. https://doi.org/10.3390/sym17081172