1 Introduction

Harmonic (or wave) maps [4] are critical points of the Dirichlet’s energy

$$\begin{aligned} \mathcal{E }(\varphi )=\frac{1}{2}\int \limits _M \vert \text{ d}\varphi \vert ^2 \text{ vol}_g \end{aligned}$$

where \(\varphi : (M, g) \rightarrow (N, h)\) denotes a smooth map between two (semi-) Riemannian manifolds \(M\) and \(N\). Harmonic morphisms form a subclass of harmonic maps characterized by an additional condition of transversal conformality. Both notions enjoy an extensive study and a deep mathematical understanding, cf. [13]. Among their various generalizations, we distinguish the \(r\)-harmonic maps and morphisms [8], defined in a similar way with respect to the energy \(\mathcal E _r(\varphi )=\tfrac{1}{r}\int _M \vert \text{ d}\varphi \vert ^r \text{ vol}_g, r>0\). These developments make harmonic maps and morphisms a very appealing tool in applications, as shown by the successfully concept of sigma model in theoretical physics. In [15], we showed that they can also play a rôle in general relativity. This rôle is due to the local duality between \(r\)-harmonic morphisms with one-dimensional timelike fibers defined on a space-time and shear-free relativistic perfect fluids \((U, p = \frac{1}{3}(r - 3)\lambda ^{r}, \rho = \lambda ^{r})\) on that space-time, where \(U\) is the unitary vector field tangent to the fibers and \(\lambda \) is the dilation of the harmonic morphism [15, Prop. 3.2]. This duality is known to geometers through the notion of foliations that produce (\(r\)-)harmonic morphisms [11].

Harmonic morphisms (and, to a lesser extent, \(r\)-harmonic morphisms) with one dimensional fibres have been classified under various curvature restrictions on the domain metric, see e.g. [9, 12] and for a review [1, 13]. In this note we want to point out that the local part of Pantilie’s classification result [12] for harmonic morphisms on Riemannian four-manifolds satisfying Einstein (vacuum) condition extends to Lorentzian four-manifolds solving the Einstein field equations (with fluid source). In the dual perspective of relativistic fluids, the counterpart of this result is the positive answer to the following

Conjecture ([16]). In general relativity, if the velocity vector field of a barotropic perfect fluid (\(p + \rho \ne 0\) and \(\rho = \rho (p)\)) is shear free, then either the expansion or the rotation of the fluid vanishes.

in the case when \(\rho =-3p\). This is already known to be true [7, 18], but in the present approach, the proof is simpler and is written in coordinate-free differential geometric language and can provide a new insight into this conjecture still not established despite a long history of confirmations in particular cases.

For a recent discussion about the conjecture and the progress that have been made, see [5, 19].

In the following section, we review the basic definitions and results needed thereafter; in Sect. 3, we give the proof of the main result, and we end with the proofs of some simple technical Lemmas.

Throughout the paper, \((M,g)\) will be a Lorentzian manifold and \(\mathrm Ric \), \(\mathrm Scal \) will denote its Ricci and scalar curvature, respectively. All the considerations are local.

2 Basic material

2.1 Perfect fluids coupled with gravity

Definition 1

[10, 17] Let \((M,g)\) be a four-dimensional space-time. A triple \((U, p, \rho )\) is called (relativistic) perfect fluid if

  1. (i)

    \(U\) is a timelike future-pointing unit vector field on \(M\), called the flow vector field,

  2. (ii)

    \(\rho , p: M\rightarrow \mathbb{R }\) are the real functions called mass (energy) density and pressure, respectively,

  3. (iii)

    the stress-energy tensor of the fluid is conserved:

    $$\begin{aligned} \mathrm div \left(p \, g + (p + \rho )U^{\flat } \otimes U^{\flat }\right)=0. \end{aligned}$$

If instead of \((iii)\), we impose the stronger condition

\((iii)^\prime \) the Einstein equations are satisfied

$$\begin{aligned} \mathrm Ric - \frac{1}{2}\mathrm Scal \cdot g = p \, g + (p + \rho )U^{\flat } \otimes U^{\flat }, \end{aligned}$$
(1)

then \((U, p, \rho )\) is called (relativistic) perfect fluid coupled with gravity.

Condition \((iii)\) of the definition decomposes into the fluid’s equations:

$$\begin{aligned} (\rho + p)\mathrm div U + U(\rho )&= 0 \quad (\text{ the} \text{ conservation} \text{ of} \text{ energy} \text{ along} \text{ the} \text{ flow}),\nonumber \\ (\rho +p)\nabla _{U}U + \mathrm grad ^{\mathcal{H }}p&= 0 \quad (\text{ the} \text{ Euler} \text{ equations}), \end{aligned}$$
(2)

where \(\mathrm grad ^{\mathcal{H }}p\) is the spatial pressure gradient, that is, the component orthogonal to \(U\). If moreover the perfect fluid is coupled with gravity, we have the following block-diagonal structure of the Ricci tensor of \((M,g)\):

$$\begin{aligned} \mathrm Ric (U,U)&= \frac{1}{2}(\rho + 3p), \nonumber \\ \mathrm Ric (X,U)&= 0, \quad \forall X \perp U, \nonumber \\ \mathrm Ric (X,Y)&= \frac{1}{2}(\rho - p)g(X,Y) , \quad \forall X,Y \perp U. \end{aligned}$$
(3)

Note that \(\mathrm Scal = \rho -3p\).

In this paper, we are concerned with perfect fluids coupled with gravity satisfying \(\rho = -3 p\). This equation of state is supposed to represent a hypothetical form of matter called quintessence characterized by negative pressure and \(-1 < p/\rho \le 0\). Moreover, as shown by Vilenkin [20], a randomly oriented distribution of (infinitely thin) straight strings averaged over all directions behaves like a perfect fluid with \(\rho =-3p\).

2.2 Harmonic morphisms and curvature restrictions

Recall [1, 6] that a harmonic morphism \(\varphi : (M,g) \rightarrow (N,h)\) between (semi-) Riemannian manifolds can be characterized as a harmonic map (i.e., \(\mathrm trace \nabla \text{ d}\varphi =0\)) which is moreover horizontally weakly conformal of dilation \(\lambda : M \rightarrow \mathbb{R }_+\) (i.e., \(\varphi ^* h = \lambda ^2 g \vert _{(\mathrm Ker \text{ d}\varphi )^\perp \times (\mathrm Ker \text{ d}\varphi )^\perp }\)).

Let \(\varphi \) be a harmonic morphism defined on a four-dimensional space-time taking values in a three-dimensional Riemannian manifold, denote by \(U\) the unit vertical timelike vector that spans \(\mathcal{V }= \mathrm Ker \text{ d}\varphi \) and by \(\mathcal{H }= \mathcal{V }^\perp \) the horizontal distribution associated with \(\varphi \). Then it is known that

$$\begin{aligned} -\nabla _U U + \mathrm grad ^{\mathcal{H }}(\ln \lambda )&= 0,\end{aligned}$$
(4)
$$\begin{aligned} \mathrm div U + 3U(\ln \lambda )&= 0,\end{aligned}$$
(5)
$$\begin{aligned} (\mathcal L _{U}g)(X,Y)-\frac{2}{3}\mathrm div U \cdot g(X,Y)&= 0, \quad \forall X,Y \in \mathcal{H }, \end{aligned}$$
(6)

where (4) is the fundamental equation of the harmonic morphism \(\varphi \), (5) is a geometric identity and (6) is equivalent to the horizontal conformality of \(\varphi \).

Recall [1, p.59] also that a section \(\sigma \) of \((\otimes ^r \mathcal{H }) \otimes (\otimes ^s \mathcal{H }^*)\) is called basic if \((\mathcal L _V \sigma )^\mathcal{H }=0\) for all \(V \in \Gamma (\mathcal{V })\). In particular, a function \(f\) on \(M\) is basic if \(V(f)=0\) and a horizontal vector field \(X\) on \(M\) is basic if \([V, X]^\mathcal{H }=0\), for all \(V \in \Gamma (\mathcal{V })\); the latter condition means that \(X\) is projectable on \(N\).

The fundamental vector field of the map \(\varphi \) is \(V := \lambda U\). Analogously to [1, Lemma 11.7.2], the importance of this vector field is given by the fact that it allows us to rewrite Eq. (4) as follows:

$$\begin{aligned}{}[V, X]=0, \quad \forall X \ \text{ basic} \text{ vector} \text{ field} \text{ on}\ M. \end{aligned}$$
(7)

Consider the 1-form dual to \(V\), defined by \(\vartheta (X) := -\lambda ^{-2}g(X, V)\), \(\forall X\), and the 2-form \(\Omega :=\text{ d}\vartheta \). We notice that \(\Omega (X,Y)=\lambda ^{-2}g([X,Y],V)\) for all \(X, Y \in \Gamma (\mathcal{H })\).

Lemma 1

Let \(\varphi :(M^4,g) \rightarrow (N^3,h)\) be a harmonic morphism defined on a four-dimensional space-time. Then,

  1. (i)

    \(i_W \Omega =0\), for all \(W \in \Gamma (\mathcal{V })\) (\(\Omega \) is a horizontal \(2\)-form)

  2. (ii)

    \(\mathcal{H }\) is integrable if and only if \(\Omega =0\) (\(\Omega \) is the integrability \(2\)-form of \(\mathcal{H }\));

  3. (iii)

    \(\mathcal L _W \Omega =0\) or, equivalently, \(W(\Omega (X,Y))=0\), for all \(X, Y\) basic vectors and for all \(W \in \Gamma (\mathcal{V })\) (\(\Omega \) is a basic 2-form).

Analogously to [12], [1, p. 343], we can prove the following

Proposition 1

Let \((M,g)\) be four-dimensional space-time, \((N,h)\) a Riemannian 3-manifold and \(\varphi :(M,g) \rightarrow (N,h)\) a harmonic morphism with one-dimensional timelike fibers tangent to the unit vector \(U\). Let \(\lambda : M \rightarrow \mathbb{R }_+\) denote the dilation of \(\varphi \). Then, we have the identities:

$$\begin{aligned} \mathrm Ric (U,U)&= \Delta \ln \lambda + 4 U(U(\ln \lambda )) - 6 [U(\ln \lambda )]^2 + \frac{\lambda ^{2}}{2} \vert \Omega \vert ^2,\end{aligned}$$
(8)
$$\begin{aligned} \mathrm Ric (X,U)&= 2 X(U(\ln \lambda )) - \frac{\lambda }{2} \{ \delta \Omega (X) + 2\Omega (X, \mathrm grad \ln \lambda )\},\end{aligned}$$
(9)
$$\begin{aligned} \mathrm Ric (X,Y)&= (\varphi ^* \mathrm Ric ^N) (X, Y) + g(X, Y) \Delta \ln \lambda - 2 X (\ln \lambda ) Y(\ln \lambda ) + \frac{\lambda ^{2}}{2} \langle i_X \Omega , i_Y \Omega \rangle ,\nonumber \\ \end{aligned}$$
(10)

where \(\mathrm Ric ^N\) is the Ricci curvature of \((N,h)\) and \(X,Y\) are horizontal vectors.

Let us recall [1, 12] also that a \(\varphi \) is of Killing type iff \(\mathrm grad ^{\mathcal{V }}\lambda = 0\), of warped product type iff it has totally geodesic fibers (i.e., \(\nabla _U U=0\)) and \(\mathcal{H }\) is integrable and of type three (T) iff \(\vert \mathrm grad ^{\mathcal{V }}\lambda \vert \) is a nonzero function of \(\lambda \). The local part of Pantilie’s classification result [12] states that if \((M,g)\) is an orientable (Rie-mannian) Einstein four-manifold, then \(\varphi \) belongs to one of these types.

2.3 The duality

In the setting of the previous subsection, from a dual perspective, Eqs. (4), (5) can be identified respectively with the Euler equations and the energy conservation along the flow (2) for the relativistic fluid \((U, p = -\frac{1}{3}\lambda ^2, \rho = \lambda ^2)\), while (6) tells that the fluid is shear free. In this case, the additional assumption (3) for the fluid to be coupled with gravity becomes

$$\begin{aligned} \mathrm Ric (U,U)&= \mathrm Ric (X,U)= 0 \nonumber \\ \mathrm Ric (X,Y)&= \frac{2}{3}\lambda ^2 g(X,Y), \quad \forall X,Y \perp U. \end{aligned}$$
(11)

Conversely, given a perfect fluid \((U, p, \rho = -3p)\), we can always integrate it locally to obtain a (local) harmonic morphism from \(M\) into some Riemannian 3-manifold \((N,h)\) such that \(\text{ d}\varphi (U)=0\) and \(\rho \) equal the squared dilation of \(\varphi \). For more details see [15].

Remark 1

If the associated harmonic morphism \(\varphi \) is of Killing type, then the fluid is expansion free (i.e., \(\mathrm div U=0\)), while if it is of warped product type, then the perfect fluid is irrotational (i.e., \(\Omega =0\)).

3 The main result

Following the same lines in [1, 12], we prove the following

Theorem 1

Let \((M,g)\) be four-dimensional space-time and \(\varphi :(M,g) \rightarrow (N,h)\) be a harmonic morphism into a Riemannian manifold, with one-dimensional timelike fibers tangent to the unit vector \(U\). Let \(\lambda \) denote its dilation. If \(\mathrm Ric (X, U)=0\) and \(\mathrm Ric (X, Y)= \nu g(X,Y)\) for all \(X, Y\) orthogonal to \(U\) and for some function \(\nu : M \rightarrow \mathbb{R }\), then locally \(\varphi \) is either of Killing type or of warped product type.

Proof

Let \((M,g)\) be a four-dimensional space-time and \(\varphi :(M, g)\rightarrow (N,h)\) be a harmonic morphism into a Riemannian manifold.

Choose a local orthogonal frame \(\{X, Y, Z\}\) of basic horizontal vector fields around a regular point \(x\) in \(M\); we may suppose that their lengths satisfy \(\vert X \vert = \vert Y \vert = \vert Z \vert = 1/\lambda \) and that \(Z\) satisfies \(i_Z\Omega = 0\) Footnote 1. Then, the contractions \(i_X\Omega \) and \(i_Y\Omega \) are both basic and orthogonal and \(\vert i_X\Omega \vert ^2=\vert i_Y\Omega \vert ^2=\lambda ^2\Omega (X,Y)^2\).

Since according to the hypothesis, \(\mathrm Ric \vert _{\mathcal{H }\times \mathcal{H }}\) must be proportional to \(g^{\mathcal{H }}\), Eq. (10) applied to all pairs of vectors from \(\{X, Y, Z\}\) gives us: \((\varphi ^* \mathrm Ric ^N) (X, Y) - 2 X (\ln \lambda ) Y(\ln \lambda )=0\), \((\varphi ^* \mathrm Ric ^N) (Y, Z) - 2 Y (\ln \lambda ) Z(\ln \lambda )=0\), and \((\varphi ^* \mathrm Ric ^N) (X, Z) - 2 X (\ln \lambda ) Z(\ln \lambda )=0\). In particular, \(X(\ln \lambda )Y(\ln \lambda )\), \(Y(\ln \lambda )Z(\ln \lambda )\), and \(Z(\ln \lambda )X (\ln \lambda )\) are all basic functions. Let \(W\) be the domain of the frame \(\{X, Y, Z, V\}\). Consider the following closed subset in \(W\):

$$\begin{aligned} S = \{x \in W : X_x(\ln \lambda ) = Y_x(\ln \lambda ) = 0\}. \end{aligned}$$

Denote its complement in \(W\) by \(S^c\)—this will be an open subset of \(M\).

Case I - on the open set \(S^c\). Suppose that \(X_x(\ln \lambda ) \ne 0\). By applying Equation (10) to the pair \(\{X + Y, X - Y\}\), we deduce that \(X(\ln \lambda )^2 - Y(\ln \lambda )^2\) is basic; moreover, since \(X(\ln \lambda )Y(\ln \lambda )\) is basic too, \(X (\ln \lambda )\) and \(Y(\ln \lambda )\) have to be both basic. Then also \(Z(\ln \lambda )\) is basic (since \(Z(\ln \lambda )X (\ln \lambda )\) is basic). Hence \(\mathrm grad (V(\ln \lambda ))\) is vertical.

Assume that at some point, \(\mathrm grad (V(\ln \lambda )) \ne 0\) so by continuity this holds in a neighborhood.

  • As \(V(\ln \lambda )\) is non-constant and its level surfaces are horizontal, then \(\mathcal{H }\) is integrable (so \(\Omega =0\)). Applying Equation (9) combined with the hypothesis \(\mathrm Ric (X,U)=0\), we obtain that \(X(U(\ln \lambda ))=0\), or equivalently \(X(\ln \lambda )V(\ln \lambda )=0\). As we are on \(S^c\), this would imply that \(V(\ln \lambda )=0\), a contradiction.

  • So \(\mathrm grad (V(\ln \lambda )) = 0\) and then \(V(\ln \lambda )=c\) a constant.

  • Assume that \(c \ne 0\).

  • Replacing \(V(\ln \lambda )=c\) in Equation (9), we obtain:

    $$\begin{aligned} 0= - 2c X(\ln \lambda ) - \frac{\lambda ^2}{2} \{ \delta \Omega (X) + 2\Omega (X, \mathrm grad \ln \lambda )\}. \end{aligned}$$
    (12)
  • Notice that \(\Omega (X, \mathrm grad \ln \lambda )\!=\!\lambda ^2 Y(\ln \lambda )\Omega (X, Y)\) and this implies \(V(\lambda ^{-2}\Omega (X, \mathrm grad \ln \lambda ))=0\). A more elaborate computation leads to: So taking the derivative of (12) along \(V\) will give

    $$\begin{aligned} 0= V(\lambda ^4) \lambda ^{-2}\{ \delta \Omega (X) + 2\Omega (X, \mathrm grad \ln \lambda )\}. \end{aligned}$$

    Therefore, the term inside the brackets must vanish; substituting this back into (12) gives us \(X(\ln \lambda )=0\), contradiction.

Lemma 2

On \(S^c\), we have \(V\left(\lambda ^{-2}\delta \Omega (X)\right)=0\).

Therefore, the constant \(c\) must be zero; this implies \(U(\ln \lambda )=0\). We have proved that on \(S^c\) (so on \(\overline{S^c}\)), \(\varphi \) is of Killing type.

Case II: on the open set \(\text{ int}S\). Let

$$\begin{aligned} A = \{x \in \mathrm int S : Z_x(\ln \lambda ) = 0\}. \end{aligned}$$
  • \((II_a)\) On the subset \(A\), we have \(\mathrm grad ^{\mathcal{H }}\ln \lambda =0\), so \(\mathrm grad \ln \lambda \) is vertical. Thus, around a point where \(\lambda \) is non-constant, its level surfaces are horizontal so \(\mathcal{H }\) is integrable and \(\varphi \) is of warped product type (if \(\lambda \) is constant, \(\mathrm div U =0\) and \(\varphi \) is of Killing type).

  • \((II_b)\) On the open subset \(A^c \cap \mathrm int S\), we have \(X(\ln \lambda ) = 0\), \(Y(\ln \lambda ) = 0\) and \(Z(\ln \lambda ) \ne 0\). The identity (10) and \(\text{ Ric}(X,X)=\mathrm Ric (Z,Z)\) (consequence of our hypothesis) imply

    $$\begin{aligned} 4 Z(\ln \lambda )^2 + \lambda ^4 \Omega (X,Y)^2 = 2\left((\varphi ^* \mathrm Ric ^N) (Z, Z)-(\varphi ^* \mathrm Ric ^N) (X, X)\right). \end{aligned}$$

Since the right-hand term is a basic function, the derivative along \(V\) of the left-hand term must vanish. This gives us immediately, by using Lemma 7:

$$\begin{aligned} -\lambda ^2 \Omega (X,Y)^2 \dfrac{V(\ln \lambda )}{Z(\ln \lambda ) } = 2\lambda ^{-2}Z(V(\ln \lambda )). \end{aligned}$$
(13)

Combine this relation with the following

Lemma 3

On \(A^c \cap \mathrm int S\), we have

$$\begin{aligned} \delta \Omega (Z)= -\lambda ^2 \Omega (X,Y)^2 \dfrac{V(\ln \lambda )}{Z(\ln \lambda ) }. \end{aligned}$$
(14)

to obtain

$$\begin{aligned} \delta \Omega (Z)= 2\lambda ^{-2}Z(V(\ln \lambda )). \end{aligned}$$
(15)

Now Equation (9) with \(\mathrm Ric (Z,U)=0\) implies

$$\begin{aligned} 4[Z(V(\ln \lambda ))-Z(\ln \lambda )V(\ln \lambda )]-\lambda ^2 \delta \Omega (Z)=0. \end{aligned}$$

Substituting (14) in the above relation, we get

$$\begin{aligned} Z(V(\ln \lambda ))-2Z(\ln \lambda )V(\ln \lambda )=0. \end{aligned}$$
(16)

Reinserting (16) in the derived constraint (13), we obtain

$$\begin{aligned} V(\ln \lambda )\left[4Z(\ln \lambda )^2 + \lambda ^4 \Omega (X,Y)^2\right]=0. \end{aligned}$$

Since \(Z(\ln \lambda )\) cannot vanish, we must have \(V(\ln \lambda )=0\) so that \(\varphi \) is of Killing type on the considered subset. \(\square \)

Taking Remark 1 into account and that coupling with gravity (11) is a special case of Theorem 1 hypothesis, we can state the following

Corollary 1

If the velocity vector field of a perfect fluid coupled with gravity satisfying \(\rho =-3p\) is shear free, then either the expansion or the rotation of the fluid vanishes.

Notice that the condition \(\mathrm Ric (U,U)=0\) from (3) was not employed.

Corollary 2

A harmonic morphism with one-dimensional timelike fibers from an Einstein space-time or from a vacuum space-time is either of Killing type or of warped product type.

The above Corollary is the Lorentzian analog of the result in [12] (see also [1, p. 380]). Note that the change of sign in the identity (10) simplifies radically the Case \((II_b)\) with respect to the Riemannian case so that type (T) morphisms are now excluded. Recall that type (T) harmonic morphisms allow both expansion and rotation (for instance on \(\mathbb{R }^4\) with the Eguchi–Hanson-type metric [14]).

Remark 2

The relativistic fluids with \(\rho = -3p\) (and harmonic morphisms) represent a genuine particular case of the shear-free conjecture. For an equation of state \(\rho = w\, p\) with \(w \ne -3\) (for \(r\)-harmonic morphisms with \(r\ne 2\)) in the Ricci identity (10), it will appear a second-order term in \(\mathrm Hess (\ln \lambda )\), as we can check by performing a conformal change in metric to render the \(r\)-harmonic morphism a harmonic morphism. This will make inapplicable the entire argument used in the present case.

4 The Proofs of Lemmas

Proof of the Lemma 2

The basic ingredients are Lemma 1 and the following identity [1, p. 119]

$$\begin{aligned} \left(\nabla ^{\varphi }_{X}\text{ d}\varphi (Y) \right)^{\widehat{}} - \nabla _X^\mathcal{H } Y = X(\ln \lambda )Y + Y(\ln \lambda )X - g(X,Y) \mathrm grad ^\mathcal{H } \ln \lambda , \end{aligned}$$

where \(\widehat{}\) stands for the horizontal lift operator. Notice that the first term in the above identity is a basic vector field. With this in hand, we can check by straightforward but lengthy calculation that

$$\begin{aligned} \delta \Omega (X)&= \lambda ^{2}\big [2Y(\ln \lambda )\Omega (X, Y)+Y(\Omega (X,Y)) + \Omega \left(\left(\nabla _Y^{\varphi }\text{ d}\varphi (Y)\right)^{\widehat{}}+ \left(\nabla _Z^{\varphi }\text{ d}\varphi (Z)\right)^{\widehat{}},X\right)\\&\quad +\Omega \left(Y, \left(\nabla _Y^{\varphi }\text{ d}\varphi (X)\right)^{\widehat{}}\,\,\right)\big ] \end{aligned}$$

and, since each term inside the brackets is basic, the conclusion follows. \(\square \)

Proof of the Lemma 3

An easy computation shows us that

$$\begin{aligned} \delta \Omega (Z)= -\lambda ^4 \Omega (X,Y) g([X,Y],Z). \end{aligned}$$
(17)

But \([X,Y](\ln \lambda )=0\) on \(\mathrm int S\) implies

$$\begin{aligned}{}[X,Y]^{\mathcal{H }}(\ln \lambda ) - \Omega (X,Y) V(\ln \lambda )=0, \end{aligned}$$

where we have used \(\Omega (X,Y)=\lambda ^{-2}g([X,Y],V)\).

As on \(A^c \cap \mathrm int S\), we have \([X,Y]^{\mathcal{H }}(\ln \lambda )=\lambda ^{2}g([X,Y],Z)Z(\ln \lambda )\), it follows that

$$\begin{aligned} g([X,Y],Z)= \Omega (X,Y) \dfrac{V(\ln \lambda )}{\lambda ^2 Z(\ln \lambda ) }. \end{aligned}$$

Reinserting in (15) gives us the result. \(\square \)