Text Vslie
Text Vslie
cohomology
Andrei Mikhailov
Abstract
We show that the BRST cohomology of the massless sector of the
Type IIB superstring on AdS5 × S 5 can be described as the relative
cohomology of an infinite-dimensional Lie superalgebra. We explain
how the vertex operators of ghost number 1, which correspond to
conserved currents, are described in this language. We also give some
algebraic description of the ghost number 2 vertices, which appears
to be new. We use this algebraic description to clarify the structure
of the zero mode sector of the ghost number two states in flat space,
and initiate the study of the vertices of the higher ghost number.
6 Conclusion 36
A Exactness of (77) 37
1
1 Introduction
Pure spinor formalism [Ber00] is a generalization of the BRST formalism
with the ghost fields constrained to satisfy a nonlinear (quadratic) equation:
λα Γm β
αβ λ = 0 (1)
where Γmαβ are the Dirac’s Gamma-matrices. A natural question arizes, what
kind of nonlinear constraints can ghost fields satisfy in a physical theory?
What if we replace (1) by an arbitrary set of equations:
λα Cαβ
i
λβ = 0 , i∈I ? (2)
Of course, this would generally speaking have nothing to do with the string
theory. But the question is, besides coming from superstring theory, what
m
special properties of Cαβ = Γmαβ are important for physics? This would be
useful to know, for example when thinking about possible generalizations of
the pure spinor formalism.
It turns out that there is some special property of (1) which plays an im-
portant role in the string worldsheet theory. This is the so-called Koszulity
— see [GKR06] and references therein. The formalism of Koszul duality was
extensively used in the study of the algebraic properties of the supersymmet-
ric Yang-Mills theories in [MS04b, MS04a], and in the classification of the
possible deformations of these theories in [MS09].
In this paper we will study the BRST cohomology of the massless sector of
the Type IIB superstring in AdS5 × S 5 . We will use the formalism of Koszul
duality to gain better understanding of the massless BRST cohomology.
The BRST cohomology counts infinitesimal deformations of the back-
ground AdS5 ×S 5 , also called “linearized excitations” or “gravitational waves”.
From the point of view of the string worldsheet theory, they are identified
with the massless vertex operators. Understanding the properties of these
vertex operators is important already because of their role in the scatter-
ing theory. Indeed, the correlation function of vertex operators is the main
ingredient in the string theory computation of the S-matrix.
Main results
1. We show that the cohomology of the BRST complex of the Type IIB
SUGRA on AdS5 × S 5 is equivalent to some relative Lie algebra coho-
mology.
2
2. We classify the vertex operators of the ghost number 1, which corre-
spond to the densities of the local conserved charges
4. We show that there are nontrivial cohomology classes at the ghost num-
ber three; we conjecture that these nontrivial cohomology classes ob-
struct the nonphysical ghost number two vertices of [BBMR11, Mik12]
∂m An + ∂n Am = 0 (4)
3
to satisfy certain system of equations. We have shown in [Mik12] that these
equations become incompatible for the unphysical vertices. The one-loop
properties of the sigma-model in AdS5 × S 5 were studied in [Val02, Ber05b,
MV09].
If we do not discuss loop effects, then the dilaton coupling does not mat-
ter. At the purely classical level, the theory deformed by a nonphysical
operator is consistent. However, such nonphysical deformations are proba-
bly inconsistent with those physical deformations which break symmetries.
We can also say that “nonphysical deformations and physical deformations,
generally speaking, obstruct each other”. In other words, nonphysical in-
finitesimal deformations only exist in those backgrounds which have global
symmetries, such as flat space and AdS5 × S 5 [Mik14]. The nonphysical
states should be absent in the semirelative cohomology, i.e. if we request
that the cochains be annihilated by the b0 − b̄0 . See [Mik14] for the analysis
of vertices in flat space.
4
coordinates xm and 16 fermionic coordinates θα . This is the supersymmetric
space-time, we will call it M :
M = R10|16 (5)
The basic superfield is the vector potential Aα (x, θ). For every α ∈ {1, . . . , 16},
the corresponding Aα is a scalar function:
Aα : M → R (6)
The equations of motion of the theory are encoded in the following construc-
tion. Let us consider the “covariant derivatives”:
∂ β ∂
∇α = α
+ Γm
αβ θ + Aα (x, θ) (7)
∂θ ∂xm
It turns out [Nil81, Wit86] that the equations of motion of SUSY Maxwell
theory are equivalent to the constraint:
• There exists a differential operator ∇m = ∂
∂xm
+ Am (x, θ) such that:
{∇α , ∇β } = Γm
αβ ∇m (8)
The nontrivial requirement of the constraint is that the LHS of (8) is pro-
portional to Γm
αβ , because the most general structure would be:
m1 m2 m3 m4 m5
Γm
αβ ∇m + Γαβ Xm1 m2 m3 m4 m5 (9)
where Xm1 ...m5 = X(x, θ)m1 ...m5 some function on the superspace. Equiva-
lently, the constraint (8) can be written:
Γαβ
m1 m2 m3 m4 m5 {∇α , ∇β } = 0 (10)
With the constraint (10) satisfied, we consider (8) as the definition of ∇m .
The pure spinor interpretation of (10) is due to [How91].
5
3 Lie algebra cohomology and solutions of
the SUSY Maxwell theory
3.1 Vacuum solution
(0)
Let us consider the vacuum solution Aα (x, θ) = 0. In this case ∇α = ∇α =
∂ β ∂
∂θα
+ Γm
αβ θ ∂xm . The vacuum solution is invariant under the supersymmetry
algebra susy generated by the operators Sα :
∂ β ∂
Sα = α − Γm αβ θ (11)
∂θ ∂xm
(0)
We observe that {Sα , ∇α } = 0, and in this sense the vacuum solution is
susy-invariant. It turns out that the operators ∇(0) themselves generate the
same (isomorphic) algebra susy as do Sα . This can be explained using the
interpretation of M as the coset space of susy. Let us consider the abstract
algebra susy generated by todd even
α and tm with the commutation relations:
m even
{todd odd
α , tβ } = Γαβ tm (12)
and other commutators all zero. Let us interpret xm and θα as coordinates
on the group manifold of the corresponding Lie group:
g = exp(θα todd m even
α + x tm ) (13)
Then ∇α acts as the multiplication by tα on the left, and Sα as the multipli-
odd
cation by todd
α on the right. We can consider the universal enveloping algebra
U susy as a representation of susy, by the left multiplication. Then the
regular representation can be considered as its dual, which will be denoted
(U susy)′ .
6
3.2 Deformations of solutions and cohomology
The deformation of the given solution Aα (x, θ) is:
Aα 7→ Aα + δAα (16)
where δAα should satisfy:
{∇α , δAβ } = Γm
αβ δAm (17)
The fact that the LHS is proportional to Γm αβ is a nontrivial constraint on
δAβ , and if it is satisfied than (17) becomes the definition of δAm .
Let us introduce pure spinors λα satisfying:
λα Γm β
αβ λ = 0 (18)
Using these pure spinors, Eq. (17) can be written:
Qv = 0 (19)
where Q = λα ∇α (20)
and v = λα δAα (21)
Therefore the problem of classifying the infinitesimal deformations of the
vacuum solution is reduced to the computation of the cohomology of Q.
7
∞
P n is a commutative algebra with quadratic relations. This
L
Notice that
n=0
algebra is Koszul dual to the universal enveloping of a Lie algebra U L.
Brief review of (23) The Koszul duality implies that the following se-
quence:
. . . −→ HomC (P 2 , U L) −→ HomC (P 1 , U L) −→ U L −→ C −→ 0 (24)
is exact, and therefore provides a free resolution of the U L-module C. This
fact depends on special properties of the quadratic constraint (1).
In (24) the action of U L on U L is by the left multiplication, and the
action of the differential involves the right multiplication by the ∇α :
dϕ(p) = ϕ(λα p)∇α (25)
Here on the right hand side we have the product of ∇α ∈ U L with ϕ(λα p) ∈
U L. In other words, for ϕ ∈ HomC (P n , U L) we have:
dϕ = µright α
U L (∇α ) ◦ ϕ ◦ µP (λ ) (26)
where µP (λα ) : P n → P n+1 is a multiplication of a polinomial by λα ∈ P 1 ,
and µright
U L (∇α ) is the right multiplication by ∇α in U L. (The composition
ϕ ◦ µ(λα ) is of the type P n → U L; we then multiply by ∇α ∈ U L.)
Since we have a projective resolution of C, we can now use it to compute
the Lie algebra cohomology of L with coefficients in V , i.e. ExtU L (C, V ). It
is the cohomology of the following sequence:
0 −→ HomU L (U L, V ) −→ HomU L (HomC (P 1 , U L), V ) −→ . . . (27)
n n+1
. . . −→ HomU L (HomC (P , U L), V ) −→ HomU L (HomC (P , U L), V ) −→ . . .
where the differential is induced by (26) and acts as follows. For f ∈
HomU L (HomC (P n , U L), V ), the df ∈ HomU L (HomC (P n+1 , U L), V ) is evalu-
ated on ϕ ∈ HomC (P n+1 , U L) as follows:
(df )(ϕ : P n+1 → U L) = f (µright α
U L (∇α ) ◦ ϕ ◦ µP (λ )) (28)
There is an isomorphism:
P n ⊗C V ≃ HomU L (HomC (P n , U L), V ) (29)
p ⊗ v 7→ [ϕ 7→ ϕ(p)v] (30)
Here “ϕ(p)v” means the action of ϕ(p) ∈ U L on the element v of the repre-
sentation V of U L. This isomorphism relates (27) to (22).
8
Special case The cohomology problem described in Section 3.2 corre-
sponds to the particular case of V = (U susy)′ . As we have just explained,
this is equivalent to the computation of the Lie algebra cohomology:
H • (L, (U susy)′ ) (31)
Notice that susy = L/I and therefore (U susy)′ is naturally a representation
of L, by the left multiplication. To calculate this cohomology, we notice that
the following complex:
. . . −→ U L ⊗C Λ2 I −→ U L ⊗C I −→ U L −→ U susy −→ 0 (32)
is a free resolution of U susy as a U L-module. This means that:
H n (QBRST ; (U susy)′ ) = H n (L, (U susy)′ ) = H n (I, C) (33)
(cf. Corollary 69 of [MS04a]). More specifically, the ghost number one vertex
operator λα δAα corresponds to the first cohomology:
′
1 I
H (I, C) = (34)
[I, I]
I
This has the following physical interpretation. The space [I,I] can be iden-
tified with the space of field strengths. Then (34) tells us that the classical
solutions are linear functionals on the space of field strengths. Indeed, given
a classical solution, we can compute the value of the field strenght on this
classical solution. Therefore, the space of classical solutions is expected to
be dual to the space of field strengths, as we indeed observe in (34).
I
Explicit description of [I,I] Elements W α of I were introduced in Eq.
(15). Consider the projection of W α to I/[I, I], i.e. W α mod [I, I]. We
conjecture that all the other elements of I/[I, I] can be obtained from W α
by commuting with ∇α , i.e. acting with susy. This means that all the gauge
invariant operators at the linearized level are W α and its derivatives.
9
4.1 BRST complex
The BRST complex of Type IIB SUGRA in AdS5 ×S 5 [BH02, BC01, Ber05b]
is based on the coset space G/G0 where G is the Lie supergroup corresponding
to the Lie superalgebra g = psu(2, 2|4) and G0 is the subgroup corresponding
to g0̄ = so(1, 4) ⊕ so(5). A Z4 -grading of g plays an important role. The
generators of g are denoted:
The subalgebra g0̄ is generated by t0[mn] , g3̄ by t3α , g1̄ by t1α̇ , and g2̄ by t2m .
The index [mn] of t0[mn] runs over a union of two sets: the set of choices of
2 different elements m, n from {0, . . . 4}, and the set of choices of 2 different
elements m, n from {5, . . . , 9}. This corresponds to the split of g0̄ into the
direct sum of so(1, 4) and so(5). Both t3α and t1α̇ transform as spinors of both
so(1, 4) and so(5) under the adjoint action of g0̄ , and t2m transform as vectors.
The BRST complex computing supergravity excitations on the back-
ground AdS5 × S 5 is:
Q Q QBRST
BRST
. . . −→ Homg0̄ (U g , P n ) −→
BRST
Homg0̄ U g , P n+1 −→ . . . (36)
where f•• • are the structure constants of g, and QBRST is given by:
QBRST = QLBRST + QR
BRST (38)
where QLBRST = λαL L(t3α ) (39)
and QRBRST = λα̇R L(t1α̇ ) (40)
Here L(t) is the left multiplication by t. We will use the notation P p,q
for theLspace of polynomials of the order p in λL and q in λR . Therefore
P n = p+q=n P p,q .
10
More generally, we can consider the cohomology with coefficients in an
arbitrary representation V of g:
Q Q Q
BRST
. . . −→ V ⊗g0̄ P n −→
BRST
V ⊗g0̄ P n+1 −→
BRST
... (41)
E2p,q = H p (QR q L
BRST ; H (QBRST ; V )) (42)
∇Lα , ∇R 0
α̇ , t[mn] (43)
3
Frobenius reciprocity implies a relation between (36) and (41), see [Mik11b].
11
where the indices α, α̇ and [mn] run over the same sets as in (35), and with
the following relations:
where Eqs. (44) and (45) are the definitions of ∇Lm and ∇R
m . The coefficients
f•• • are the structure constants of psu(2, 2|4) in the basis (35). We will
introduce the following notation for this Lie algebra:
deg(∇Lα ) = 1
deg(∇R
α̇ ) = − 1 (51)
H • Ltot ; g0̄ ; V
(52)
4
For introduction into the Lie algebra cohomology, see [Kna88, FF88]
12
We claim that this cohomology coincides with the BRST cohomology:
ξ1 ∧ ξ2 ∧ · · · ∧ ξq 7→ c(ξ1 ∧ ξ2 ∧ . . . ∧ ξq ) (55)
13
Elements of the linear space Ltot /g0̄ are, by definition in Section 4.2, nested
commutators of ∇L s plus nested commutators of ∇R s. We define Rc as the
following function of λL and λR :
⊗q
Rc(λL , λR ) = c λαL ∇Lα + λα̇R ∇R
α̇ (56)
for c ∈ C q (Ltot ; g0̄ ; V )
We observe that:
1. The action of QLie on grp C p+q (Ltot ; g0̄ ; V ) coincides with the ac-
[H • (LL ,V )] [H • (LR ,C)]
on pr=0 C q+r (LL ; V ) ⊗g0
L
tion of the operator QLie + QLie
grp C p−r (LR ; C)
14
4.4 An analogue of the Koszul resolution
In fact, it is possible to glue two Koszul resolutions (one for LL and another
for LR ) along g0̄ , as we will now explain5 . Similarly to (24), consider the
following BRST-type complex:
(notations as in (24)), and Homg0̄ means linear maps invariant under the
following action of g0̄ :
We will call the two terms on the right hand side of (63) dL ϕ and dR ϕ. We
will introduce the abbreviated notation for the terms of (62):
0 −→ C −→ X 0 −→ X 1 −→ . . . (65)
There is a bigrading: X n = p+q=n X p,q where X p,q = Homg0̄ (U Ltot , P p,q );
L
Proof Being (U Ltot , U g0̄ )-injective follows from Section 1 of [Hoc56] (Lemma
1). Note that every term of (62) is a direct sum of finite-dimensional represen-
tations of g0̄ . This implies that the kernel and the image of every differential
is a direct g0̄ -submodule as required in [Hoc56]. It remains to prove the ex-
actness. We will prove the equivalent statement, that the cohomology of the
truncated complex:
0 −→ X 0 −→ X 1 −→ . . . (66)
is only nonzero in the zeroth term: H 0 = C. We will use the spectral
sequence of the bicomplex d = dL + dR . Let us first calculate the cohomology
5
Note in the revised version: we are greatful to the referee of [CMV13] for pointing out
an error in the original version of this subsection
15
of dL . We will “normal order” the elements of U Ltot by putting elements of
U LR to the left and elements of U LL to the right. This gives an isomorphism
of linear spaces:
Homg0̄ U Ltot , P n−p, p = HomC U LL , PLn−p ⊗ HomC U LR , PRp
(67)
The differential dL only acts on the HomC U L , PL , while HomC U LR , PRp
n−p
L
Corollary This means that for any U Ltot -module W , the Ext(U Ltot ,U g0̄ ) (W, C)
can be computed as the cohomology of the following complex:
. . . −→ HomU Ltot W , Homg0̄ (U Ltot , P n ) −→
16
As in Section 3.3, there is an isomorphism of complexes (73) and (41):
Homg0̄ U g ⊗g0̄ A , P •
(79)
17
Geometrically, this is the space of A′ -valued functions fa (g, λ3 , λ1 ) where the
index a enumerates a basis of A′ , such that for h ∈ G0̄ :
More precisely, this is the space of Taylor series of sections of the pure
spinor bundle over AdS5 × S 5 ; the universal enveloping algebra is the space
of finite linear combinations, i.e. we do not care about the convergence of the
Taylor series f . Equation (80) says that f is a section of a bundle over the
homogeneous space. On the other hand, Eq. (81) requires that f transform
in a fixed representation A′ under the group G0̄ of global rotations around
g = 1.
The space of Taylor series, as a representation of the global rotations G0̄ ,
is the direct sum of infinitely many finite-dimensional representations:
M
Homg0̄ (U g , P • ) = A ⊗ Homg0̄ U g ⊗g0̄ A , P •
(82)
A
18
Special notations for summation over repeating indices. As already
introduced in (35), the index m enumerates the basis of the vector repre-
sentation of g0̄ = so(1, 4) ⊕ so(5), and runs from 0 to 9; more precisely,
m ∈ {0, . . . , 4} enumerates vectors of so(1, 4), and m ∈ {5, . . . , 9} vectors of
so(5). For a vector v m we denote:
m
m v if m ∈ {0, . . . , 1}
v = (85)
−v m if m ∈ {5, . . . , 9}
I
Proposition. As a representation of g, [I,I]
is generated by the following
objects6 :
Tm2 = ∇Lm − ∇R
m (87)
0
T[mn] = [∇Lm , ∇Ln ] − [∇R R
m , ∇n ] (88)
1
ZαL = ∇Lα − [ ∇Lm , [∇Lm , ∇Lα ] ] (89)
10
1
Zα̇R = ∇Rα̇ − [ ∇R , [∇R R
m , ∇α̇ ] ] (90)
10 m
Notice that [(∇Lm − ∇R L R
m ) , (∇n − ∇n )] ∈ [I, I] implies that:
Similarly, [(∇Lm − ∇R L L L
m ), [(∇m − ∇m ), ∇α ]] ∈ [I, I] implies that:
1 mα̇ R R
∇Lα − fα [∇m , ∇α̇ ] = −ZαL mod [I, I] (92)
10
We will write “≡ 0” instead of “= 0 mod [I, I]”.
6 1
The coefficient 10 depends on the choice of normalization for ∇α ; in our conventions
fαβ = Γαβ , and the projection pr(∇m ) of ∇m to g satisfies: (adpr(∇m ) )2 |g3̄ = 1 — no
m m
summation over m.
19
The (30|32)-dimensional linear space generated by Tm2 , T[mn]0
, ZαL , Zα̇R is
closed under the action of g. It must be the adjoint representation of g.
For example, let us consider {∇Lα , ZβL }. Modulo [I, I] this is same as
{[∇R R L
m , ∇α̇ ] , Zβ }, and using (46), (47) and (48) this is proportional to Tm .
Suppose that the coefficients C are such that this expression belongs to I. We
will prove that it also belongs to J, using the induction in q — the number
of commutators. Suppose that for q < n, all such expressions lie in J. We
will prove that for q = n, (93) is also in J.
Notice that:
X
α1 ...α5 L L L L 1 mβ R R
C [∇α1 , {∇α2 , . . . {∇αq−1 , ∇αq − fαq [∇m , ∇β̇ ] }] . . .}] ∈ J
10
α
⃗
(94)
L 1 mβ R R
because ∇α − 10 fα [∇m , ∇β̇ ] ∈ J. Therefore, it remains to prove that the
following expression belongs to J:
X
C α1 ...α5 [∇Lα1 , {∇Lα2 , . . . {∇Lαq−1 , fαq mβ [∇R R
m , ∇β̇ ]}]}] (95)
α
⃗
20
That the RHS of (96) is proportional to fmα α̇ and the RHS of (97) is pro-
portional to fmα̇ α follows from (44) and (45).
To calculate {∇Lα , Zα̇R }, {∇R L 2
α̇ , Zα } and [∇m , Tn ] we start with the
following observation:
0 ≡ {∇Lα − ∇R L R L L R R 0
α , ∇α̇ − ∇α̇ } = {∇α , ∇α̇ } + {∇α , ∇α̇ } − 2tαα̇ (99)
Also notice:
{[∇Lm , ∇Lβ̇ ] , ∇R L β L R L
α̇ − ∇α̇ } ≡ fmβ̇ {∇β , ∇α̇ − ∇α̇ } =
= [∇Lm , {∇Lβ̇ , ∇R L L L R L
α̇ − ∇α̇ }] − {∇β̇ , [∇m , ∇α̇ − ∇α̇ ]} ≡
This implies:
Similarly:
fmβ β̇ {∇R
β̇
, ZαL } − fmα γ̇ {∇Lβ , Zγ̇R } = −fαβ n [∇R R L
m , ∇n − ∇ n ] (102)
Taking into account (98), we get the following system of equations for Xαα̇ =
{∇Lα , Zα̇R } and Xmn = [∇Lm , ∇Ln − ∇Rn ]:
We have to prove that there are no other solutions. Let us use the identity:
n
fm αβ fαβ n = 16 δm (107)
21
Contracting (103) and (104) with f α̇β̇ k and f αβ k we get:
This implies:
fmα̇ γ fk α̇β̇ Xγ β̇ + fkα̇ γ fm α̇β̇ Xγ β̇ = 0 (110)
Let us assume that the pair (m, k) is such that:
either m ∈ {0, . . . , 4} and k ∈ {5, . . . , 9}
or m ∈ {5, . . . , 9} and k ∈ {0, . . . , 4};
then (110) implies that for such pairs (m, k) the expression fmα̇ γ fk α̇β̇ Xγ β̇ is
symmetric under the exchange m ↔ k. But Xmk is always antisymmetric
under such an exchange. Therefore Eq. (108) implies that Xmk is only
nonzero when either both m and k belong to {0, . . . , 4}, or both m and k
belong to {5, . . . , 9}. This means that Xmk is proportional to fmk • , and we
0
can define T[pq] from (105). Then (103) gives:
2fm(α̇| γ Xγ|β̇) − fγ|β̇) [pq] Y[pq] = 0 (111)
I
To summarize, [I,I]
is a finite-dimensional space, the adjoint representa-
tion of g.
7
This is the Poincaré duality, Section VI.3 of [Kna88].
22
where xi and yi are elements of I, with the equivalence relations:
a ≃ a + [x, y] ∧ z + [y, z] ∧ x + [z, x] ∧ y (114)
We do not have the complete analysis at the ghost number two. It must be
true that H2 (I) correspond to the space of gauge-invariant8 operators at a
marked point in AdS5 × S 5 . This is an infinite-dimensional representation of
g. The simplest element of H2 (I) is:
O = C αα̇ (∇Lα − WαR ) ∧ (∇R L
α̇ − Wα̇ ) (115)
where C αα̇ is the coefficient in the decomposition of the quadratic Casimir
operator of psl(4|4):
C = C mn t2m t2n + C [mn][pq] t0[mn] t0[pq] + C αα̇ t3α t1α̇ + C α̇α t1α̇ t3α (116)
This should correspond to the value of the dilaton. It should be possible to
obtain other fields by acting on (115) with ∇Lα and ∇Rα̇ .
23
5.1 Ghost number 1.
I
The space [I,I]
is generated by ∇Lm − ∇R L L α α̇
m , [∇m , ∇n ], WL and WR . We
observe:
[∇Lm , ∇Ln ] = − [∇R R
m , ∇n ] mod [I, I] (119)
As a representation of susy, this space should be the dual to susy+Lorentz.
We observe:
γ 1 n
{∇(α , Γm L
β)γ WL } ≡ Γαβ [∇n , ∇m ]
L
(120)
2
As explained in [Maf09], Eq. (120) implies that ∇α WLγ is proportional to
(Γmn )γα [∇Ln , ∇Lm ].
≡ 4 [[∇L[k , ∇Ll ] , ∇R L R L
m − ∇m ] ∧ (∇n] − ∇n] ) = 0 (126)
24
— here [[∇L[k , ∇Ll ], ∇Lm] ] = 0 because of the Jacobi identity. To prove (125)
we observe that when calculating ∇j ϕ for any element ϕ of H2 (I), we can use
either ∇Lj ϕ or ∇R j ϕ. Since both terms on the right hand side of (122) are in
H2 (I), we are free to use ∇Lj when calculating ∇j ([∇Lk , ∇Ll ] ∧ [∇R R
m , ∇n ]) and
∇R L L R R
j when calculating ∇j ([∇m , ∇n ] ∧ [∇k , ∇l ]). Those are both zero because
of the Jacobi identity.
∇n ∇k Hklm = 0 (128)
The dilaton The difference ALm − AR m should be identified with the first
derivative of the dilaton ∂m ϕ. Notice that:
4 L L
∇n (ALm − AR R R
m ) = [∇k , ∇(m ] ∧ [∇n) , ∇k ] (131)
3
This is in agreement with the statement that (122) is a linear combination
of the Riemann-Christoffel tensor Rklmn and the derivatives of the dilaton
9
Since the homology of I is I-invariant, we can calculate either ∇L k
n ∇ Hklm or
∇R k
n ∇ Hklm ;
R k
it is easier to calculate ∇n ∇ Hklm
25
∂[l gk][m ∂n] ϕ. Indeed, we have:
3
g lm [∇Lk , ∇Ll ] ∧ [∇R R L L R R L R
m , ∇n ] + [∇m , ∇n ] ∧ [∇k , ∇l ] − ∇n (Ak − Ak ) = 0
4
(132)
which is the Einstein’s equation Rkn = 0 for the Ricci tensor Rkn = g lm Rklmn ,
if we identify:
∇l (ALm + AR L R
m ) = ∇[l (Am] + Am] ) (135)
∇k ∇l (ALm + AR
m) = 0 (136)
26
A state on which ALm +AR m is nonzero is described in [Mik12]. It is obtained
as the flat space limit of the nonphysical AdS vertex of [BBMR11] with the
internal commutator taking values in g2̄ (using the notations of Section 4.1).
In this case ALm +AR m is constant — the gradient of the “asymmetric dilaton”.
Besides being constant, ALm + AR m can also be depending on x linearly. To
L R
obtain the state with Am + Am depending linearly on x, we have to consider
the flat space limit of the nonphysical vertex Bab j a ∧ j b with the internal
commutator f ab c Bab taking values in g0̄ [BBMR11, Mik12]. It depends on a
constant antisymmetric tensor Bmn . The leading term in the flat space limit
is a trivial constant NSNS B-field Bmn dxm ∧ dxn , which can be gauged away.
Discarding the terms with θ’s, the leading nontrivial term is:
4 4
!
X X
Bmn dxm ∧ xn (dxk xk ) − dxn (xk xk ) (137)
k=0 k=0
This does not solve the SUGRA equations ∂ n Hnml = 0, instead ∂ n Hnml is
proportional to Bmn dxm ∧ dxn — a constant 2-form.
In terms of the unintegrated vertex, the observable ALm + AR m should be
identified as follows. It is proportional to ∂ n Bmn in the gauge where the
vertex has ghost number (1, 1), i.e. only λL λR terms, no λL λL and λR λR
terms10 .
27
The gradient of the dilaton corresponds to A− +
n , while An does not have a clear
interpretation in the Type IIB supergravity. The “observable” A+ n is dual to
the unphysical vertex of [Mik12]. The unphysical vertex is not BRST trivial.
However, as we explained in [Mik12], it should be thrown away because it
leads to a quantum anomaly in the worldsheet sigma-model at the 1-loop
level.
O = xL ∧ xR (145)
28
We suspect that the cohomology at the ghost numbers 3 and 4 is a finite-
dimensional space, and is in some way related to the unphysical states of
[BBMR11, Mik12].
We will start by proving the vanishing theorem for the super-Maxwell co-
homology at the ghost number higher than 1. We will then point out that the
SUGRA BRST complex is amlost the tensor product of two super-Maxwell
complexes (the “left sector” and the “right sector”). If it were, literally, the
tensor product, that would indeed imply the vanishing theorem at the ghost
number > 2. But in fact, even in flat space there is some “interaction” be-
tween the left and the right sector, and this leads to a nontrivial cohomology
at least at the ghost number 3.
Sketch of the proof This fact is well-known in the pure spinor formal-
ism. At the ghost number 0, the cohomology is formed by the constants (no
dependence on λ, x and θ). At the ghost number 1, the cohomology is the
solutions of the free Maxwell equation and the free Dirac equation. The van-
ishing of the cohomology at the ghost number 2 is equivalent to the following
two statements: 1) for any current jm such that ∂m jm = 0 always exists the
gauge field Fmn satisfying ∂[k Flm] = 0 and ∂m Fmn = jn and 2) for any spinor
ψ exists a spinor ϕ such that Γm ∂m ϕ = ψ. The vanishing of the cohomology
at the ghost number 3 is equivalent to the statement that for any ρ exists
jm such that ∂m jm = ρ. All these facts are proven in any graduate course of
classical electrodynamics.
29
The cohomology of (149) is the tensor product of the cohomologies of two
super-Maxwell complexes. Therefore it is only nontrivial at the ghost num-
bers 0,1 and 2. However, in the Type IIB BRST complex there is no sepa-
ration of x into xL and xR . The actual BRST complex is therefore different
from (149):
α ∂ m β ∂ α̂ ∂ m β̂ ∂
QSUGRA = λL + Γαβ θL m + λR + Γα̂β̂ θR m (150)
∂θLα ∂x ∂θRα̂ ∂x
The difference is that the left and the right sector have a common x instead
of separate xL and xR . We also write:
QSUGRA = QL + QR (151)
where QL and QR are the first and second terms on the right hand side of
(150).
Vanishing theorem: HQn SUGRA = 0 for n > 4. Let us consider, for example,
a vertex of the ghost number 5.
30
were every ϕj is linear in λL . We observe that all these ϕj s are annihilated
by QL (because QL V = 0 and QL does not act on θR ):
QL ϕj = 0 (153)
We also observe that in the leading term, the coefficient of ϕk is annihilated
by λR ∂θ∂R . This implies:
k+1
V = QSUGRA λ3R θR
ϕk (λL , θL , x) +
4 k+1 k+2
+ λR θR ϕk+1 (λL , θL , x) + θR ϕ̃k+2 (λL , θL , x) + . . . (154)
This means that we are able to increase the order of the leading term by
adding a QSUGRA -exact expression. The induction in k proves the Theo-
rem.
But is it true that HQn SUGRA = 0 for n = 3 and n = 4? It turns out that at
least for n = 3 the cohomology is nontrivial. The fact that the cohomology
at the ghost number higher than 2 is nontrivial is (for us) unexpected. We
will leave this for future research, giving here only an example.
+ (θL Γp λL ) θL Γp xm
L Γ m f [λ R θR
4
] + gn [λ L θ 4 n
]x
L R F̂ Γq θR (λR Γq θR )
(156)
4
where f [λR θR ] is chosen so that:
∂ l ∂
λR + (θR Γ λR ) l xnR Γn F̂ Γq θR (λR Γq θR ) + f [λR θR
4
] =0 (157)
∂θR ∂xR
and g[λL θL4 ] is chosen so that:
∂ l ∂
(θL Γp λL )θL Γp (xm n n 4
λL + (θL Γ λL ) l L Γm Γn − 10x L ) + g [λL θ L ] =0
∂θL ∂xL
(158)
31
4
Such f [λR θR ] and g n [λL θL4 ] exist because the expression xnR Γn F̂ satisfies the
“right” Dirac equation:
∂ n
x Γn F̂ Γk = 0 (159)
∂xkR R
and the expression (xm n
L Γm Γn − 10xL ) satisfies the “left” Dirac equation:
∂
Γk (xm n
L Γm Γn − 10xL ) = 0 (160)
∂xkL
We will now prove that Φ[F ] depends on xL and xR only in the combination
xL + xR . Indeed, for a constant cm let us introduce Ξ[c, F ] as follows:
m ∂ ∂
Ξ[c, F ] = c − m Ψ[F ] =
∂xmL ∂xR
= (θL Γp λL ) θL Γp cm Γm xnR Γn F̂ Γq θR (λR Γq θR ) +
+ (θL Γp λL ) θL Γp cm Γm f [λR θR 4
] −
− (θL Γp λL ) θL Γp (xm L Γm c n
Γn − 10(x L c))F̂ Γq θR (λR Γq θR ) −
− g n [λL θL4 ]cn F̂ Γq θR (λR Γp θR ) (161)
32
on the cohomology of QSMaxw⊗SMaxw . The answer to the question (163) is
positive only if Ξ[c, F ] is a coboundary in this complex. The cohomology of
QSMaxw⊗SMaxw is the tensor product of two super-Maxwell solutions. We will
now prove that Ξ[c, F ] represents a nonzero element of:
1 m ∂ ∂
H c − m , SMaxw(xL ) ⊗ SMaxw(xR ) (164)
∂xm L ∂xR
∂ α′ β̇
Γm
αα′ ψ (xL , xR ) = 0 (165)
∂xm
L
∂ αβ̇ ′
ψ (xL , xR )Γm
β̇ ′ β̇
=0 (166)
∂xm
R
where hat over letter stands for the contraction with the gamma-matrices,
e.g. x̂R = Γm xm
R .Let us analize the possibility of (167) being in the image
m ∂ ∂
of c ∂xm
− ∂xm :
L R
αβ̇
?
ĉx̂R F̂ − (x̂L ĉ − 10(xL · c))F̂ =
? m ∂ ∂ αβ̇ m n αβ̇ m n αβ̇ m n
=c − ϕ x
mn L L x + χ x x
mn L R + σ x
mn R Rx (168)
∂xm L ∂xm R
33
′
The left Dirac equation on χ implies Γm α β̇
αα′ χmn = 0, therefore:
β̇
α′ β̇ n
10 x̂R F̂ = −2Γm
αα′ σmn xR (170)
α
Equations (172) and (173) imply that the traces of σ and s are zero:
αβ̇ αβ̇
σmm = smm =0 (174)
Consider the cohomology in the sector Dirac ⊗ Dirac, and more specifically
those elements of it which have linear x-dependence. It turns out that this
cohomology is identified with the quadratic in x solutions f of the “double
34
Dirac equation” modulo solutions presentable as a sum of a solution of the
left Dirac equation and a solution of the right Dirac equation:
∂ m ∂ n α′ α̇′
Γ ′ Γ ′f (x) = 0
∂xm αα ∂xn α̇α̇
but ∄ s and σ such that: f αα̇ = sαα̇ + σ αα̇ (176)
∂ m α′ α̇ ∂ αα̇′ n
m
Γαα′ s = 0 and σ Γα̇′ α̇ = 0
∂x ∂xn
Indeed, given such an f αα̇ with the quadratic x-dependence, we construct
ψ(c) in the following way:
∂
ψ(c) = ĉΓn f (xR ) + ξ(xL , c) (177)
∂xnR
where ξ is some solution of the left Dirac equation, chosen so that QLie ψ =
0; such a solution always exists because H 2 (R10 , Dirac) = 0. Suppose
that ψ is in the image of QLie acting on the quadratic (in xL|R ) elements of
Dirac ⊗ Dirac, i.e.:
? m ∂ ∂
ψ(c) = c − m (σ⟨xR ⊗ xR ⟩ + χ⟨xR ⊗ xL ⟩ + ϕ⟨xL ⊗ xL ⟩) (178)
∂xmL ∂xR
The part of ψ(c) linear in xR would be:
∂ m ∂
−cm m σ⟨x⊗2 R ⟩+c χ⟨xR ⊗ xL ⟩ (179)
∂xR ∂xm
L
This implies:
∂ ∂ ∂
Γm m ψ(c)⟨xR ⟩ = 10Γn n f (xR ) = −Γm m σ⟨x⊗2 R ⟩ (180)
∂c ∂xR ∂xR
in other words f = s + σ where σ satisfies the right Dirac equation and s the
left Dirac equation. This contradicts (176).
Eq. (171) has f αα̇ = ||x||2 F̂ αα̇ with a 5-form F̂ ; there are also solutions
corresonding to a 3-form or 7-form Ĝ:
1
f = Ĝ||x||2 − x̂Γp ĜΓp x̂ (181)
52
and a 1-form or 9-form Â:
1
f = Â||x||2 − x̂Γp ÂΓp x̂ (182)
28
This means that the cohomology at the ghost number 3 at least includes
states with the quantum number of a bispinor.
35
5.3.3 Dual picture
We conjecture that the dual element of H3 (I) is of the form:
1 ′
− WLα ∧ WRβ̇ (Γmn )β̇β̇ ′ ∧ [∇R R
m , ∇n ] +
2
1 ′
+ WLα (Γmn )αα′ ∧ WRβ̇ ∧ [∇Lm , ∇Ln ] (183)
2
6 Conclusion
In this paper we presented a relation between the cohomology of the pure
spinor BRST complex in AdS space and the relative Lie algebra cohomology.
We used this relation to develop a “dual” point of view on the vertex
operators in Type IIB. In this approach, instead of looking at the vertex
operators, we look at the dual linear space which is identified with the gauge-
invariant local operators of the Type IIB SUGRA. This works both in flat
space and in AdS. We observe that some elements of the BRST cohomology
36
do not correspond to any physical states, e.g. the A+ of (140). It turns
out that there are also vertex operators at the ghost number three. They
correspond to the obstructions for nonlinear deformations in the actions.
Physically, these obstructions should not be present.
Such “unphysical” elements should go away if we restrict the BRST com-
plex to the operators annihilated by the Virasoro constraints. We do not
know what this restriction means from the point of view of the Lie algebra
cohomology.
We conclude that the BRST complex (36) in AdS5 × S 5 and its flat space
limit (118) both have rich mathematical structure. But at the same time
the cohomology does not give a complete description of the supergravity
excitations. The difference is in some unphysical states. These unphysical
states have polynomial x-dependence, as opposed to the usually considered
exponential x-dependence. This polynomial (or “zero-momentum”) sector
could be important in the calculation of the scattering amplitude, because the
momentum conservation implies that the product of the scattering vertices
has zero total momentum.
A Exactness of (77)
This is similar to the proof of the exactness of the standard Koszul resolution
of the Lie algebra in [Kna88]. For any Lie algebra L, the universal enveloping
U L is filtered so that grp U L = F p U L/F p−1 U L = S p L. The differential in
our complex acts in such a way, that we can consistently define:
0 −→ A −→ A −→ 0 (185)
37
On the other hand, the factor-complex X p /X p−1 is:
. . . −→ S p−2 Ltot /g0̄ ⊗C Λ2 I ⊗C A −→ S p−1 Ltot /g0̄ ⊗C I ⊗C A −→
(186)
This is exact, being the de Rham complex of the linear space I times functions
of additional “inert” variables corresponding to a complement to g0̄ + I in
Ltot . By induction, the complexes X p are exact for all values of p, and
therefore the complex (77) is exact.
Acknowledgments
I would like to thank Nathan Berkovits for discussions and the anonymous
referee for useful suggestions. This work was supported in part by the Min-
istry of Education and Science of the Russian Federation under the project
14.740.11.0347 “Integrable and algebro-geometric structures in string the-
ory and quantum field theory”, and in part by the RFFI grant 10-02-01315
“String theory and integrable systems”.
References
[BBMR11] O. A. Bedoya, L. I. Bevilaqua, A. Mikhailov, and V. O. Riv-
elles, Notes on beta-deformations of the pure spinor super-
string in AdS(5) x S(5), Nucl.Phys. B848 (2011), 155–215,
arXiv/1005.0049.
[BC01] N. Berkovits and O. Chandia, Superstring vertex operators in an
ads(5) x s(5) background, Nucl. Phys. B596 (2001), 185–196,
hep-th/0009168.
[Ber00] N. Berkovits, Super-Poincare covariant quantization of the su-
perstring, JHEP 04 (2000), 018, hep-th/0001035.
[Ber05a] , BRST cohomology and nonlocal conserved charges,
JHEP 02 (2005), 060, hep-th/0409159.
[Ber05b] , Quantum consistency of the superstring in AdS(5) x S(5)
background, JHEP 03 (2005), 041, hep-th/0411170.
38
[BH02] N. Berkovits and P. S. Howe, Ten-dimensional supergravity con-
straints from the pure spinor formalism for the superstring, Nucl.
Phys. B635 (2002), 75–105, hep-th/0112160.
[Mik13] , A generalization of the Lax pair for the pure spinor su-
perstring in AdS5 x S5, arXiv/1303.2090.
39
[MS04a] M. Movshev and A. S. Schwarz, Algebraic structure of Yang-Mills
theory, arXiv/hep-th/0404183.
40