You are currently browsing the monthly archive for March 2014.
Let be an irreducible polynomial in three variables. As
is not algebraically closed, the zero set
can split into various components of dimension between
and
. For instance, if
, the zero set
is a line; more interestingly, if
, then
is the union of a line and a surface (or the product of an acnodal cubic curve with a line). We will assume that the
-dimensional component
is non-empty, thus defining a real surface in
. In particular, this hypothesis implies that
is not just irreducible over
, but is in fact absolutely irreducible (i.e. irreducible over
), since otherwise one could use the complex factorisation of
to contain
inside the intersection
of the complex zero locus of complex polynomial
and its complex conjugate, with
having no common factor, forcing
to be at most one-dimensional. (For instance, in the case
, one can take
.) Among other things, this makes
a Zariski-dense subset of
, thus any polynomial identity which holds true at every point of
, also holds true on all of
. This allows us to easily use tools from algebraic geometry in this real setting, even though the reals are not quite algebraically closed.
The surface is said to be ruled if, for a Zariski open dense set of points
, there exists a line
through
for some non-zero
which is completely contained in
, thus
for all . Also, a point
is said to be a flecnode if there exists a line
through
for some non-zero
which is tangent to
to third order, in the sense that
for . Clearly, if
is a ruled surface, then a Zariski open dense set of points on
are a flecnode. We then have the remarkable theorem (discovered first by Monge, and then later by Cayley and Salmon) asserting the converse:
Theorem 1 (Monge-Cayley-Salmon theorem) Let
be an irreducible polynomial with
non-empty. Suppose that a Zariski dense set of points in
are flecnodes. Then
is a ruled surface.
Among other things, this theorem was used in the celebrated result of Guth and Katz that almost solved the Erdos distance problem in two dimensions, as discussed in this previous blog post. Vanishing to third order is necessary: observe that in a surface of negative curvature, such as the saddle , every point on the surface is tangent to second order to a line (the line in the direction for which the second fundamental form vanishes). This surface happens to be ruled, but a generic perturbation of this surface (e.g.
) will no longer be ruled, although it is still negative curvature near the origin.
The original proof of the Monge-Cayley-Salmon theorem is not easily accessible and not written in modern language. A modern proof of this theorem (together with substantial generalisations, for instance to higher dimensions) is given by Landsberg; the proof uses the machinery of modern algebraic geometry. The purpose of this post is to record an alternate proof of the Monge-Cayley-Salmon theorem based on classical differential geometry (in particular, the notion of torsion of a curve) and basic ODE methods (in particular, Gronwall’s inequality and the Picard existence theorem). The idea is to “integrate” the lines indicated by the flecnode to produce smooth curves
on the surface
; one then uses the vanishing (1) and some basic calculus to conclude that these curves have zero torsion and are thus planar curves. Some further manipulation using (1) (now just to second order instead of third) then shows that these curves are in fact straight lines, giving the ruling on the surface.
Update: Janos Kollar has informed me that the above theorem was essentially known to Monge in 1809; see his recent arXiv note for more details.
I thank Larry Guth and Micha Sharir for conversations leading to this post.
A core foundation of the subject now known as arithmetic combinatorics (and particularly the subfield of additive combinatorics) are the elementary sum set estimates (sometimes known as “Ruzsa calculus”) that relate the cardinality of various sum sets
and difference sets
as well as iterated sumsets such as ,
, and so forth. Here,
are finite non-empty subsets of some additive group
(classically one took
or
, but nowadays one usually considers more general additive groups). Some basic estimates in this vein are the following:
Lemma 1 (Ruzsa covering lemma) Let
be finite non-empty subsets of
. Then
may be covered by at most
translates of
.
Proof: Consider a maximal set of disjoint translates of
by elements
. These translates have cardinality
, are disjoint, and lie in
, so there are at most
of them. By maximality, for any
,
must intersect at least one of the selected
, thus
, and the claim follows.
Lemma 2 (Ruzsa triangle inequality) Let
be finite non-empty subsets of
. Then
.
Proof: Consider the addition map from
to
. Every element
of
has a preimage
of this map of cardinality at least
, thanks to the obvious identity
for each
. Since
has cardinality
, the claim follows.
Such estimates (which are covered, incidentally, in Section 2 of my book with Van Vu) are particularly useful for controlling finite sets of small doubling, in the sense that
for some bounded
. (There are deeper theorems, most notably Freiman’s theorem, which give more control than what elementary Ruzsa calculus does, however the known bounds in the latter theorem are worse than polynomial in
(although it is conjectured otherwise), whereas the elementary estimates are almost all polynomial in
.)
However, there are some settings in which the standard sum set estimates are not quite applicable. One such setting is the continuous setting, where one is dealing with bounded open sets in an additive Lie group (e.g. or a torus
) rather than a finite setting. Here, one can largely replicate the discrete sum set estimates by working with a Haar measure in place of cardinality; this is the approach taken for instance in this paper of mine. However, there is another setting, which one might dub the “discretised” setting (as opposed to the “discrete” setting or “continuous” setting), in which the sets
remain finite (or at least discretisable to be finite), but for which there is a certain amount of “roundoff error” coming from the discretisation. As a typical example (working now in a non-commutative multiplicative setting rather than an additive one), consider the orthogonal group
of orthogonal
matrices, and let
be the matrices obtained by starting with all of the orthogonal matrice in
and rounding each coefficient of each matrix in this set to the nearest multiple of
, for some small
. This forms a finite set (whose cardinality grows as
like a certain negative power of
). In the limit
, the set
is not a set of small doubling in the discrete sense. However,
is still close to
in a metric sense, being contained in the
-neighbourhood of
. Another key example comes from graphs
of maps
from a subset
of one additive group
to another
. If
is “approximately additive” in the sense that for all
,
is close to
in some metric, then
might not have small doubling in the discrete sense (because
could take a large number of values), but could be considered a set of small doubling in a discretised sense.
One would like to have a sum set (or product set) theory that can handle these cases, particularly in “high-dimensional” settings in which the standard methods of passing back and forth between continuous, discrete, or discretised settings behave poorly from a quantitative point of view due to the exponentially large doubling constant of balls. One way to do this is to impose a translation invariant metric on the underlying group
(reverting back to additive notation), and replace the notion of cardinality by that of metric entropy. There are a number of almost equivalent ways to define this concept:
Definition 3 Let
be a metric space, let
be a subset of
, and let
be a radius.
- The packing number
is the largest number of points
one can pack inside
such that the balls
are disjoint.
- The internal covering number
is the fewest number of points
such that the balls
cover
.
- The external covering number
is the fewest number of points
such that the balls
cover
.
- The metric entropy
is the largest number of points
one can find in
that are
-separated, thus
for all
.
It is an easy exercise to verify the inequalities
for any , and that
is non-increasing in
and non-decreasing in
for the three choices
(but monotonicity in
can fail for
!). It turns out that the external covering number
is slightly more convenient than the other notions of metric entropy, so we will abbreviate
. The cardinality
can be viewed as the limit of the entropies
as
.
If we have the bounded doubling property that is covered by
translates of
for each
, and one has a Haar measure
on
which assigns a positive finite mass to each ball, then any of the above entropies
is comparable to
, as can be seen by simple volume packing arguments. Thus in the bounded doubling setting one can usually use the measure-theoretic sum set theory to derive entropy-theoretic sumset bounds (see e.g. this paper of mine for an example of this). However, it turns out that even in the absence of bounded doubling, one still has an entropy analogue of most of the elementary sum set theory, except that one has to accept some degradation in the radius parameter
by some absolute constant. Such losses can be acceptable in applications in which the underlying sets
are largely “transverse” to the balls
, so that the
-entropy of
is largely independent of
; this is a situation which arises in particular in the case of graphs
discussed above, if one works with “vertical” metrics whose balls extend primarily in the vertical direction. (I hope to present a specific application of this type here in the near future.)
Henceforth we work in an additive group equipped with a translation-invariant metric
. (One can also generalise things slightly by allowing the metric to attain the values
or
, without changing much of the analysis below.) By the Heine-Borel theorem, any precompact set
will have finite entropy
for any
. We now have analogues of the two basic Ruzsa lemmas above:
Lemma 4 (Ruzsa covering lemma) Let
be precompact non-empty subsets of
, and let
. Then
may be covered by at most
translates of
.
Proof: Let be a maximal set of points such that the sets
are all disjoint. Then the sets
are disjoint in
and have entropy
, and furthermore any ball of radius
can intersect at most one of the
. We conclude that
, so
. If
, then
must intersect one of the
, so
, and the claim follows.
Lemma 5 (Ruzsa triangle inequality) Let
be precompact non-empty subsets of
, and let
. Then
.
Proof: Consider the addition map from
to
. The domain
may be covered by
product balls
. Every element
of
has a preimage
of this map which projects to a translate of
, and thus must meet at least
of these product balls. However, if two elements of
are separated by a distance of at least
, then no product ball can intersect both preimages. We thus see that
, and the claim follows.
Below the fold we will record some further metric entropy analogues of sum set estimates (basically redoing much of Chapter 2 of my book with Van Vu). Unfortunately there does not seem to be a direct way to abstractly deduce metric entropy results from their sum set analogues (basically due to the failure of a certain strong version of Freiman’s theorem, as discussed in this previous post); nevertheless, the proofs of the discrete arguments are elementary enough that they can be modified with a small amount of effort to handle the entropy case. (In fact, there should be a very general model-theoretic framework in which both the discrete and entropy arguments can be processed in a unified manner; see this paper of Hrushovski for one such framework.)
It is also likely that many of the arguments here extend to the non-commutative setting, but for simplicity we will not pursue such generalisations here.
As in the previous post, all computations here are at the formal level only.
In the previous blog post, the Euler equations for inviscid incompressible fluid flow were interpreted in a Lagrangian fashion, and then Noether’s theorem invoked to derive the known conservation laws for these equations. In a bit more detail: starting with Lagrangian space and Eulerian space
, we let
be the space of volume-preserving, orientation-preserving maps
from Lagrangian space to Eulerian space. Given a curve
, we can define the Lagrangian velocity field
as the time derivative of
, and the Eulerian velocity field
. The volume-preserving nature of
ensures that
is a divergence-free vector field:
If we formally define the functional
then one can show that the critical points of this functional (with appropriate boundary conditions) obey the Euler equations
for some pressure field . As discussed in the previous post, the time translation symmetry of this functional yields conservation of the Hamiltonian
the rigid motion symmetries of Eulerian space give conservation of the total momentum
and total angular momentum
and the diffeomorphism symmetries of Lagrangian space give conservation of circulation
for any closed loop in
, or equivalently pointwise conservation of the Lagrangian vorticity
, where
is the
-form associated with the vector field
using the Euclidean metric
on
, with
denoting pullback by
.
It turns out that one can generalise the above calculations. Given any self-adjoint operator on divergence-free vector fields
, we can define the functional
as we shall see below the fold, critical points of this functional (with appropriate boundary conditions) obey the generalised Euler equations
for some pressure field , where
in coordinates is
with the usual summation conventions. (When
,
, and this term can be absorbed into the pressure
, and we recover the usual Euler equations.) Time translation symmetry then gives conservation of the Hamiltonian
If the operator commutes with rigid motions on
, then we have conservation of total momentum
and total angular momentum
and the diffeomorphism symmetries of Lagrangian space give conservation of circulation
or pointwise conservation of the Lagrangian vorticity . These applications of Noether’s theorem proceed exactly as the previous post; we leave the details to the interested reader.
One particular special case of interest arises in two dimensions , when
is the inverse derivative
. The vorticity
is a
-form, which in the two-dimensional setting may be identified with a scalar. In coordinates, if we write
, then
Since is also divergence-free, we may therefore write
where the stream function is given by the formula
If we take the curl of the generalised Euler equation (2), we obtain (after some computation) the surface quasi-geostrophic equation
This equation has strong analogies with the three-dimensional incompressible Euler equations, and can be viewed as a simplified model for that system; see this paper of Constantin, Majda, and Tabak for details.
Now we can specialise the general conservation laws derived previously to this setting. The conserved Hamiltonian is
(a law previously observed for this equation in the abovementioned paper of Constantin, Majda, and Tabak). As commutes with rigid motions, we also have (formally, at least) conservation of momentum
(which up to trivial transformations is also expressible in impulse form as , after integration by parts), and conservation of angular momentum
(which up to trivial transformations is ). Finally, diffeomorphism invariance gives pointwise conservation of Lagrangian vorticity
, thus
is transported by the flow (which is also evident from (3). In particular, all integrals of the form
for a fixed function
are conserved by the flow.
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.
Recent Comments