You are currently browsing the tag archive for the ‘Euler-Arnold equation’ tag.
These lecture notes are a continuation of the 254A lecture notes from the previous quarter.
We consider the Euler equations for incompressible fluid flow on a Euclidean space ; we will label
as the “Eulerian space”
(or “Euclidean space”, or “physical space”) to distinguish it from the “Lagrangian space”
(or “labels space”) that we will introduce shortly (but the reader is free to also ignore the
or
subscripts if he or she wishes). Elements of Eulerian space
will be referred to by symbols such as
, we use
to denote Lebesgue measure on
and we will use
for the
coordinates of
, and use indices such as
to index these coordinates (with the usual summation conventions), for instance
denotes partial differentiation along the
coordinate. (We use superscripts for coordinates
instead of subscripts
to be compatible with some differential geometry notation that we will use shortly; in particular, when using the summation notation, we will now be matching subscripts with superscripts for the pair of indices being summed.)
In Eulerian coordinates, the Euler equations read
where is the velocity field and
is the pressure field. These are functions of time
and on the spatial location variable
. We will refer to the coordinates
as Eulerian coordinates. However, if one reviews the physical derivation of the Euler equations from 254A Notes 0, before one takes the continuum limit, the fundamental unknowns were not the velocity field
or the pressure field
, but rather the trajectories
, which can be thought of as a single function
from the coordinates
(where
is a time and
is an element of the label set
) to
. The relationship between the trajectories
and the velocity field was given by the informal relationship
We will refer to the coordinates as (discrete) Lagrangian coordinates for describing the fluid.
In view of this, it is natural to ask whether there is an alternate way to formulate the continuum limit of incompressible inviscid fluids, by using a continuous version of the Lagrangian coordinates, rather than Eulerian coordinates. This is indeed the case. Suppose for instance one has a smooth solution
to the Euler equations on a spacetime slab
in Eulerian coordinates; assume furthermore that the velocity field
is uniformly bounded. We introduce another copy
of
, which we call Lagrangian space or labels space; we use symbols such as
to refer to elements of this space,
to denote Lebesgue measure on
, and
to refer to the
coordinates of
. We use indices such as
to index these coordinates, thus for instance
denotes partial differentiation along the
coordinate. We will use summation conventions for both the Eulerian coordinates
and the Lagrangian coordinates
, with an index being summed if it appears as both a subscript and a superscript in the same term. While
and
are of course isomorphic, we will try to refrain from identifying them, except perhaps at the initial time
in order to fix the initialisation of Lagrangian coordinates.
Given a smooth and bounded velocity field , define a trajectory map for this velocity to be any smooth map
that obeys the ODE
in view of (2), this describes the trajectory (in ) of a particle labeled by an element
of
. From the Picard existence theorem and the hypothesis that
is smooth and bounded, such a map exists and is unique as long as one specifies the initial location
assigned to each label
. Traditionally, one chooses the initial condition
for , so that we label each particle by its initial location at time
; we are also free to specify other initial conditions for the trajectory map if we please. Indeed, we have the freedom to “permute” the labels
by an arbitrary diffeomorphism: if
is a trajectory map, and
is any diffeomorphism (a smooth map whose inverse exists and is also smooth), then the map
is also a trajectory map, albeit one with different initial conditions
.
Despite the popularity of the initial condition (4), we will try to keep conceptually separate the Eulerian space from the Lagrangian space
, as they play different physical roles in the interpretation of the fluid; for instance, while the Euclidean metric
is an important feature of Eulerian space
, it is not a geometrically natural structure to use in Lagrangian space
. We have the following more general version of Exercise 8 from 254A Notes 2:
Exercise 1 Let
be smooth and bounded.
- If
is a smooth map, show that there exists a unique smooth trajectory map
with initial condition
for all
.
- Show that if
is a diffeomorphism and
, then the map
is also a diffeomorphism.
Remark 2 The first of the Euler equations (1) can now be written in the form
which can be viewed as a continuous limit of Newton’s first law
.
Call a diffeomorphism (oriented) volume preserving if one has the equation
for all , where the total differential
is the
matrix with entries
for
and
, where
are the components of
. (If one wishes, one can also view
as a linear transformation from the tangent space
of Lagrangian space at
to the tangent space
of Eulerian space at
.) Equivalently,
is orientation preserving and one has a Jacobian-free change of variables formula
for all , which is in turn equivalent to
having the same Lebesgue measure as
for any measurable set
.
The divergence-free condition then can be nicely expressed in terms of volume-preserving properties of the trajectory maps
, in a manner which confirms the interpretation of this condition as an incompressibility condition on the fluid:
Lemma 3 Let
be smooth and bounded, let
be a volume-preserving diffeomorphism, and let
be the trajectory map. Then the following are equivalent:
on
.
is volume-preserving for all
.
Proof: Since is orientation-preserving, we see from continuity that
is also orientation-preserving. Suppose that
is also volume-preserving, then for any
we have the conservation law
for all . Differentiating in time using the chain rule and (3) we conclude that
for all , and hence by change of variables
which by integration by parts gives
for all and
, so
is divergence-free.
To prove the converse implication, it is convenient to introduce the labels map , defined by setting
to be the inverse of the diffeomorphism
, thus
for all . By the implicit function theorem,
is smooth, and by differentiating the above equation in time using (3) we see that
where is the usual material derivative
acting on functions on . If
is divergence-free, we have from integration by parts that
for any test function . In particular, for any
, we can calculate
and hence
for any . Since
is volume-preserving, so is
, thus
Thus is volume-preserving, and hence
is also.
Exercise 4 Let
be a continuously differentiable map from the time interval
to the general linear group
of invertible
matrices. Establish Jacobi’s formula
and use this and (6) to give an alternate proof of Lemma 3 that does not involve any integration in space.
Remark 5 One can view the use of Lagrangian coordinates as an extension of the method of characteristics. Indeed, from the chain rule we see that for any smooth function
of Eulerian spacetime, one has
and hence any transport equation that in Eulerian coordinates takes the form
for smooth functions
of Eulerian spacetime is equivalent to the ODE
where
are the smooth functions of Lagrangian spacetime defined by
In this set of notes we recall some basic differential geometry notation, particularly with regards to pullbacks and Lie derivatives of differential forms and other tensor fields on manifolds such as and
, and explore how the Euler equations look in this notation. Our discussion will be entirely formal in nature; we will assume that all functions have enough smoothness and decay at infinity to justify the relevant calculations. (It is possible to work rigorously in Lagrangian coordinates – see for instance the work of Ebin and Marsden – but we will not do so here.) As a general rule, Lagrangian coordinates tend to be somewhat less convenient to use than Eulerian coordinates for establishing the basic analytic properties of the Euler equations, such as local existence, uniqueness, and continuous dependence on the data; however, they are quite good at clarifying the more algebraic properties of these equations, such as conservation laws and the variational nature of the equations. It may well be that in the future we will be able to use the Lagrangian formalism more effectively on the analytic side of the subject also.
Remark 6 One can also write the Navier-Stokes equations in Lagrangian coordinates, but the equations are not expressed in a favourable form in these coordinates, as the Laplacian
appearing in the viscosity term becomes replaced with a time-varying Laplace-Beltrami operator. As such, we will not discuss the Lagrangian coordinate formulation of Navier-Stokes here.
Throughout this post, we will work only at the formal level of analysis, ignoring issues of convergence of integrals, justifying differentiation under the integral sign, and so forth. (Rigorous justification of the conservation laws and other identities arising from the formal manipulations below can usually be established in an a posteriori fashion once the identities are in hand, without the need to rigorously justify the manipulations used to come up with these identities).
It is a remarkable fact in the theory of differential equations that many of the ordinary and partial differential equations that are of interest (particularly in geometric PDE, or PDE arising from mathematical physics) admit a variational formulation; thus, a collection of one or more fields on a domain
taking values in a space
will solve the differential equation of interest if and only if
is a critical point to the functional
involving the fields and their first derivatives
, where the Lagrangian
is a function on the vector bundle
over
consisting of triples
with
,
, and
a linear transformation; we also usually keep the boundary data of
fixed in case
has a non-trivial boundary, although we will ignore these issues here. (We also ignore the possibility of having additional constraints imposed on
and
, which require the machinery of Lagrange multipliers to deal with, but which will only serve as a distraction for the current discussion.) It is common to use local coordinates to parameterise
as
and
as
, in which case
can be viewed locally as a function on
.
Example 1 (Geodesic flow) Take
and
to be a Riemannian manifold, which we will write locally in coordinates as
with metric
for
. A geodesic
is then a critical point (keeping
fixed) of the energy functional
or in coordinates (ignoring coordinate patch issues, and using the usual summation conventions)
As discussed in this previous post, both the Euler equations for rigid body motion, and the Euler equations for incompressible inviscid flow, can be interpreted as geodesic flow (though in the latter case, one has to work really formally, as the manifold
is now infinite dimensional).
More generally, if
is itself a Riemannian manifold, which we write locally in coordinates as
with metric
for
, then a harmonic map
is a critical point of the energy functional
or in coordinates (again ignoring coordinate patch issues)
If we replace the Riemannian manifold
by a Lorentzian manifold, such as Minkowski space
, then the notion of a harmonic map is replaced by that of a wave map, which generalises the scalar wave equation (which corresponds to the case
).
Example 2 (
-particle interactions) Take
and
; then a function
can be interpreted as a collection of
trajectories
in space, which we give a physical interpretation as the trajectories of
particles. If we assign each particle a positive mass
, and also introduce a potential energy function
, then it turns out that Newton’s laws of motion
in this context (with the force
on the
particle being given by the conservative force
) are equivalent to the trajectories
being a critical point of the action functional
Formally, if is a critical point of a functional
, this means that
whenever is a (smooth) deformation with
(and with
respecting whatever boundary conditions are appropriate). Interchanging the derivative and integral, we (formally, at least) arrive at
Write for the infinitesimal deformation of
. By the chain rule,
can be expressed in terms of
. In coordinates, we have
where we parameterise by
, and we use subscripts on
to denote partial derivatives in the various coefficients. (One can of course work in a coordinate-free manner here if one really wants to, but the notation becomes a little cumbersome due to the need to carefully split up the tangent space of
, and we will not do so here.) Thus we can view (2) as an integral identity that asserts the vanishing of a certain integral, whose integrand involves
, where
vanishes at the boundary but is otherwise unconstrained.
A general rule of thumb in PDE and calculus of variations is that whenever one has an integral identity of the form for some class of functions
that vanishes on the boundary, then there must be an associated differential identity
that justifies this integral identity through Stokes’ theorem. This rule of thumb helps explain why integration by parts is used so frequently in PDE to justify integral identities. The rule of thumb can fail when one is dealing with “global” or “cohomologically non-trivial” integral identities of a topological nature, such as the Gauss-Bonnet or Kazhdan-Warner identities, but is quite reliable for “local” or “cohomologically trivial” identities, such as those arising from calculus of variations.
In any case, if we apply this rule to (2), we expect that the integrand should be expressible as a spatial divergence. This is indeed the case:
Proposition 1 (Formal) Let
be a critical point of the functional
defined in (1). Then for any deformation
with
, we have
where
is the vector field that is expressible in coordinates as
Proof: Comparing (4) with (3), we see that the claim is equivalent to the Euler-Lagrange equation
The same computation, together with an integration by parts, shows that (2) may be rewritten as
Since is unconstrained on the interior of
, the claim (6) follows (at a formal level, at least).
Many variational problems also enjoy one-parameter continuous symmetries: given any field (not necessarily a critical point), one can place that field in a one-parameter family
with
, such that
for all ; in particular,
which can be written as (2) as before. Applying the previous rule of thumb, we thus expect another divergence identity
whenever arises from a continuous one-parameter symmetry. This expectation is indeed the case in many examples. For instance, if the spatial domain
is the Euclidean space
, and the Lagrangian (when expressed in coordinates) has no direct dependence on the spatial variable
, thus
then we obtain translation symmetries
for , where
is the standard basis for
. For a fixed
, the left-hand side of (7) then becomes
where . Another common type of symmetry is a pointwise symmetry, in which
for all , in which case (7) clearly holds with
.
If we subtract (4) from (7), we obtain the celebrated theorem of Noether linking symmetries with conservation laws:
Theorem 2 (Noether’s theorem) Suppose that
is a critical point of the functional (1), and let
be a one-parameter continuous symmetry with
. Let
be the vector field in (5), and let
be the vector field in (7). Then we have the pointwise conservation law
In particular, for one-dimensional variational problems, in which , we have the conservation law
for all
(assuming of course that
is connected and contains
).
Noether’s theorem gives a systematic way to locate conservation laws for solutions to variational problems. For instance, if and the Lagrangian has no explicit time dependence, thus
then by using the time translation symmetry , we have
as discussed previously, whereas we have , and hence by (5)
and so Noether’s theorem gives conservation of the Hamiltonian
For instance, for geodesic flow, the Hamiltonian works out to be
so we see that the speed of the geodesic is conserved over time.
For pointwise symmetries (9), vanishes, and so Noether’s theorem simplifies to
; in the one-dimensional case
, we thus see from (5) that the quantity
is conserved in time. For instance, for the -particle system in Example 2, if we have the translation invariance
for all , then we have the pointwise translation symmetry
for all ,
and some
, in which case
, and the conserved quantity (11) becomes
as was arbitrary, this establishes conservation of the total momentum
Similarly, if we have the rotation invariance
for any and
, then we have the pointwise rotation symmetry
for any skew-symmetric real matrix
, in which case
, and the conserved quantity (11) becomes
since is an arbitrary skew-symmetric matrix, this establishes conservation of the total angular momentum
Below the fold, I will describe how Noether’s theorem can be used to locate all of the conserved quantities for the Euler equations of inviscid fluid flow, discussed in this previous post, by interpreting that flow as geodesic flow in an infinite dimensional manifold.
A (smooth) Riemannian manifold is a smooth manifold without boundary, equipped with a Riemannian metric
, which assigns a length
to every tangent vector
at a point
, and more generally assigns an inner product
to every pair of tangent vectors at a point
. (We use Roman font for
here, as we will need to use
to denote group elements later in this post.) This inner product is assumed to symmetric, positive definite, and smoothly varying in
, and the length is then given in terms of the inner product by the formula
In coordinates (and also using abstract index notation), the metric can be viewed as an invertible symmetric rank
tensor
, with
One can also view the Riemannian metric as providing a (self-adjoint) identification between the tangent bundle of the manifold and the cotangent bundle
; indeed, every tangent vector
is then identified with the cotangent vector
, defined by the formula
In coordinates, .
A fundamental dynamical system on the tangent bundle (or equivalently, the cotangent bundle, using the above identification) of a Riemannian manifold is that of geodesic flow. Recall that geodesics are smooth curves that minimise the length
There is some degeneracy in this definition, because one can reparameterise the curve without affecting the length. In order to fix this degeneracy (and also because the square of the speed is a more tractable quantity analytically than the speed itself), it is better if one replaces the length with the energy
Minimising the energy of a parameterised curve turns out to be the same as minimising the length, together with an additional requirement that the speed
stay constant in time. Minimisers (and more generally, critical points) of the energy functional (holding the endpoints fixed) are known as geodesic flows. From a physical perspective, geodesic flow governs the motion of a particle that is subject to no external forces and thus moves freely, save for the constraint that it must always lie on the manifold
.
One can also view geodesic flows as a dynamical system on the tangent bundle (with the state at any time given by the position
and the velocity
) or on the cotangent bundle (with the state then given by the position
and the momentum
). With the latter perspective (sometimes referred to as cogeodesic flow), geodesic flow becomes a Hamiltonian flow, with Hamiltonian
given as
where is the inverse inner product to
, which can be defined for instance by the formula
In coordinates, geodesic flow is given by Hamilton’s equations of motion
In terms of the velocity , we can rewrite these equations as the geodesic equation
where
are the Christoffel symbols; using the Levi-Civita connection , this can be written more succinctly as
If the manifold is an embedded submanifold of a larger Euclidean space
, with the metric
on
being induced from the standard metric on
, then the geodesic flow equation can be rewritten in the equivalent form
where is now viewed as taking values in
, and
is similarly viewed as a subspace of
. This is intuitively obvious from the geometric interpretation of geodesics: if the curvature of a curve
contains components that are transverse to the manifold rather than normal to it, then it is geometrically clear that one should be able to shorten the curve by shifting it along the indicated transverse direction. It is an instructive exercise to rigorously formulate the above intuitive argument. This fact also conforms well with one’s physical intuition of geodesic flow as the motion of a free particle constrained to be in
; the normal quantity
then corresponds to the centripetal force necessary to keep the particle lying in
(otherwise it would fly off along a tangent line to
, as per Newton’s first law). The precise value of the normal vector
can be computed via the second fundamental form as
, but we will not need this formula here.
In a beautiful paper from 1966, Vladimir Arnold (who, sadly, passed away last week), observed that many basic equations in physics, including the Euler equations of motion of a rigid body, and also (by which is a priori a remarkable coincidence) the Euler equations of fluid dynamics of an inviscid incompressible fluid, can be viewed (formally, at least) as geodesic flows on a (finite or infinite dimensional) Riemannian manifold. And not just any Riemannian manifold: the manifold is a Lie group (or, to be truly pedantic, a torsor of that group), equipped with a right-invariant (or left-invariant, depending on one’s conventions) metric. In the context of rigid bodies, the Lie group is the group of rigid motions; in the context of incompressible fluids, it is the group
) of measure-preserving diffeomorphisms. The right-invariance makes the Hamiltonian mechanics of geodesic flow in this context (where it is sometimes known as the Euler-Arnold equation or the Euler-Poisson equation) quite special; it becomes (formally, at least) completely integrable, and also indicates (in principle, at least) a way to reformulate these equations in a Lax pair formulation. And indeed, many further completely integrable equations, such as the Korteweg-de Vries equation, have since been reinterpreted as Euler-Arnold flows.
From a physical perspective, this all fits well with the interpretation of geodesic flow as the free motion of a system subject only to a physical constraint, such as rigidity or incompressibility. (I do not know, though, of a similarly intuitive explanation as to why the Korteweg de Vries equation is a geodesic flow.)
One consequence of being a completely integrable system is that one has a large number of conserved quantities. In the case of the Euler equations of motion of a rigid body, the conserved quantities are the linear and angular momentum (as observed in an external reference frame, rather than the frame of the object). In the case of the two-dimensional Euler equations, the conserved quantities are the pointwise values of the vorticity (as viewed in Lagrangian coordinates, rather than Eulerian coordinates). In higher dimensions, the conserved quantity is now the (Hodge star of) the vorticity, again viewed in Lagrangian coordinates. The vorticity itself then evolves by the vorticity equation, and is subject to vortex stretching as the diffeomorphism between the initial and final state becomes increasingly sheared.
The elegant Euler-Arnold formalism is reasonably well-known in some circles (particularly in Lagrangian and symplectic dynamics, where it can be viewed as a special case of the Euler-Poincaré formalism or Lie-Poisson formalism respectively), but not in others; I for instance was only vaguely aware of it until recently, and I think that even in fluid mechanics this perspective to the subject is not always emphasised. Given the circumstances, I thought it would therefore be appropriate to present Arnold’s original 1966 paper here. (For a more modern treatment of these topics, see the books of Arnold-Khesin and Marsden-Ratiu.)
In order to avoid technical issues, I will work formally, ignoring questions of regularity or integrability, and pretending that infinite-dimensional manifolds behave in exactly the same way as their finite-dimensional counterparts. In the finite-dimensional setting, it is not difficult to make all of the formal discussion below rigorous; but the situation in infinite dimensions is substantially more delicate. (Indeed, it is a notorious open problem whether the Euler equations for incompressible fluids even forms a global continuous flow in a reasonable topology in the first place!) However, I do not want to discuss these analytic issues here; see this paper of Ebin and Marsden for a treatment of these topics.
Recent Comments