AMSI 2up
AMSI 2up
AMSI 2up
Jerry L. Kazdan
iv
CONTENTS
Contents
Chapter 1. Introduction
1. Functions of Several Variables
2. Classical Partial Differential Equations
3. Ordinary Differential Equations, a Review
1
2
3
5
11
11
11
13
18
22
29
29
29
33
36
41
43
44
59
59
60
60
64
65
67
69
75
1. INTRODUCTION
Gilbarg, D., and Trudinger, N. S., Elliptic Partial Differential Equations of Second Order, 2 nd Edition, Springer-Verlag, 1983.
Introduction
Partial Differential Equations (PDEs) arise in many applications to
physics, geometry, and more recently the world of finance. This will be
a basic course.
In real life one can find explicit solutions of very few PDEs and many
of these are infinite series whose secrets are complicated to extract. For
more than a century the goal is to understand the solutions even
though there may not be a formula for the solution.
The historic heart of the subject (and of this course) are the three fundamental linear equations: wave equation, heat equation, and Laplace
equation along with a few nonlinear equations such as the minimal surface equation and others that arise from problems in the calculus of
variations.
We seek insight and understanding rather than complicated formulas.
Prerequisites: Linear algebra, calculus of several variables, and basic
ordinary differential equations. In particular Ill assume some experience with the Stokes and divergence theorems and a bit of Fourier
analysis. Previous acquaintantance with normed linear spaces will also
be assumed. Some of these topics will be reviewed a bit as needed.
References: For this course, the most important among the following
are the texts by Strauss and Evans.
Strauss, Walter A., Partial Differential Equations: An Introduction,
New York, NY: Wiley, 1992.
John, Fritz. Partial Differential Equations, 4th ed., Series: Applied
Mathematical Sciences, New York, NY: Springer-Verlag.
Axler, S., Bourdin, P., and Ramey, W., Harmonic Function Theory,
accessible at
http://www.axler.net/HFT.pdf.
Courant, Richard, and Hilbert, David, Methods of Mathematical Physics,
vol II. Wiley-Interscience, New York, 1962.
Evans, L.C., Partial Differential Equations, American Mathematical
Society, Providence, 1998.
Jost, J., Partial Differential Equations, Series: Graduate Texts in Mathematics, Vol. 214 . 2nd ed., 2007, XIII, 356 p.
1
uxx uxy
uxy uyy
is positive definite at the origin. Must this function have its global
minimum at the origin, that is, can one conclude that u(x, y) >
u(0, 0)
for
all
(x, y)
6
=
(0, 0)?
Proof or counter example.
1. INTRODUCTION
flow away from that point and contradict the assumed equilibrium.
This is the maximum principle: if u satisfies the Laplace equation then
min u u(x, y) max u
wave equation
heat equation
Laplace equation
for (x, y)
for (x, y) .
Rt
a(x) dx
that is S 1 LS = D,
1. INTRODUCTION
so using the change of variables defined by the operator S , the differential operator L is similar to the basic operator D . Consequently
we can reduce problems concerning L to those for D .
Exercise: With Lu := Du + au as above, we seek a solution u(t), periodic with period 1 of Lu = f , assuming a(t) and f are also periodic,
a(t + 1) = a(t) etc. It will help to introduce the inner product
Z 1
g(t)h(t) dt.
hg, hi =
0
d2 u 2
+c u = 0, with c 6= 0 a constant. Before doing anything
dt2
else, we can rescale the variable t, replacing t by t/c to reduce to the
special case c = 1. Using scaling techniques can lead to deep results.
The operator Lu := u + c2 u = 0 has two types of invariance: i).
linearity in u and translation invariance in t.
Linearity in u means that
c).
L(u + v) = Lu + Lv,
L(T u) = T L(u)
sin() = b.
This means [ekt E(t)] 0. Use this to deduce that E(t) ekt E(0)
for all t 0, so the energy can grow at most exponentially].
Exercise: If a map L is translation invariant [see (1.2)], and q(t; ) :=
Let , show that q(t; ) = g(0; )et . Thus, writing Q() = q(0; ),
conclude that
Let = Q()et ,
that is, et is an eigenfunction of L with eigenvalue Q(). You find
special solutions of the homogeneous equation by finding the values of
where Q() = 0.
Exercise: Use the previous exercise to discuss the second order linear
difference equation u(x + 2) = u(x + 1) + u(x). Then apply this to find
the solution of
u(n + 2) = u(n + 1) + u(n),
n = 0, 1, 2, . . .
1. INTRODUCTION
d). Group Invariance. One can use group invariance as the key
to solving many problems. Here are some examples:
a) au + bu + cu = 0, where a, b, and c are constants. This linear equation is also invariant under translation t 7 t + , as the
example above. One seeks special solutions that incorporate the
translation invariance and then use the linearity to build the general solution.
b) at2 u + btu + cu = 0, where a, b, and c are constants. This is
invariant under the similarity t 7 t. One seeks special solutions
that incorporate the similarity invariance and then use the linearity
to build the general solution.
du
at2 + bu2
c)
= 2
, where a, b, c, and d are constants. This is
dt
ct + du2
invariant under the stretching
t 7 t,
u 7 u,
for > 0.
In each case the idea is to seek a special solution that incorporates the
u
invariance. For instance, in the last example, try v(t) = .
t
Lie began his investigation of what we now call Lie Groups by trying to
use Galois group theoretic ideas to understand differential equations.
e). Local vs Global: nonlinear. . Most of the focus above was
on local issues, say solving a differential equation du/dt = f (t, u) for
small t. A huge problem remains to understand the solutions for large
t. This leads to the qualitative theory, and requires wonderful new
ideas from topology. Note, however, that for nonlinear equations (or
linear equations with singularities), a solution might only exist for finite
t. The simplest example is
du
= u2 with initial conditions u(0) = c.
dt
The solution, obtained by separation of variables,
c
u(t) =
1 ct
blows up at t = 1/c.
f). Local vs Global: boundary value problems. Global issues
also arise if instead of solving an initial value problem one is solving a
boundary value problem such as
(1.3)
d2 u
+ a2 u = f (x) with boundary conditions u(0) = 0, u() = 0.
dx2
Here one only cares about the interval 0 x . As the following
exercise illustrates, even the case when a is a constant gives non-obvious
results.
Exercise:
a) In the special case of (1.3) where a = 0, show that a solution exists
for any f .
R
b) If a = 1, show that a solution exists if and only if 0 f (x) sin x dx =
0.
c) If 0 a < 1 is a constant, show that a solution exists for any f .
Exercise: [Maximum Principle]
a) Let u(x) be a solution of u + u = 0 for 0 < x < 1. Show that
at a point x = x0 where u has a local maximum, u cannot be
positive. If u(x0 ) = 0, what can you conclude?
b) Generalize to solutions of u + b(x)u + c(x)u = 0, assuming
c(x) > 0.
c) Say u and v both satisfy u + u = f (x) for 0 < x < 1 with
u(0) = v(0) and u(1) = v(1). Show that u(x) = v(x) for all
0 x 1.
d) Say u is a periodic solution, so u(1) = u(0) and u (1) = u(0), of
u = 1 h(x)eu
for 0 x 1,
10
12
CHAPTER 2
u
u
3
= f (x, y),
x
y
so given f (x, y) one wants u(x, y). This problem is not quite as trivial
as one might think.
u(x, 0) = h(x)
(2.1)
(x(t)) = h(t).
However, one cannot use an arbitrary curve . For an extreme example, if is vertical, that is, x(t) = const., then one cannot solve
the initial value problem (2.2) (2.5) unless h(t) const. Thus one cannot prescribe arbitrary initial data on an arbitrary curve. Even more
seriously, if one differentiates (2.6), then one finds
(2.7)
(2.2)
uy (x, y) = 0
in is
(2.3)
u(x, y) = (x),
(x(t))
dx
dh
=
,
dt
dt
13
14
u
u r u s
u
u
=
+
=a
+c
x
r x s x
r
s
u r u s
u
u
u
=
+
=b
+d
y
r y
s y
r
s
If one attempts to find a particular solution of the inhomogeneous equation uy = f in a domain , where f C (), then vertical convexity
is again needed. In fact
Proposition 2.1. One can solve uy = f for all f C ()
is vertically convex.
Proof: Just integrate.
A proof can be found in [Hormander-1, Theorems
3.5.4 and 3.7.2]. However the following argument (I
learned it from G. Schwarz) is adequate for many
domains such as the region
p in the figure. Let
f (x, y) = 1/r , where r = x2 + y 2 . Assume there
is a solution u of uy = f . Then for any > 0,
(2.8)
u(, 1) u(, 1) =
uy (, t) dt =
1
dt.
2
+ t2
Now as 0, the left side is finite but the right side becomes infinite.
This contradiction completes the proof.
3. A More General Example
a). Constant coefficient. Other first order equations can be treated
similarly. For example, the equation
(2.9)
1
1
Thus
ut + cux = 0.
It is a simple model for the following situation. Say one has water
flowing at a constant velocity c in a horizontal cylindrical pipe along
the x-axis. Initially , near x = 0 a colored dye is inserted in the water.
Ignoring possible dispersion of the dye, it will simply flow along the
pipe. The concentration u(x, t) of the dye, then is reasonably described
by the transport equation. If the initial concentration is u(x, 0) = f (x),
then by our discussion in the previous paragraph, the solution is
u(x, t) = f (x ct).
ux + 2uy = 0,
t=0
t = t2=1
c c 2cc 2c 2c
u(x,0)
u(x,1)
f(x-c)
u(x,2)===f(x)
f(x2c)
zz
15
16
t=0
2c
u(x,0) = f(x)
Exercise: Solve ux + uy = 0 with the initial value u(0, y) = 3 sin y .
Exercise: Solve ux + uy + 2u = 0 with the initial value u(0, y) = 3y .
b). Variable coefficient. First order linear equations with variable coefficients
(2.11)
u
u
+y
= f (x, y).
x
y
that is
ur =
f (r cos , r sin )
,
r
For two independent variables, the following small modification is sometimes convenient. Consider for the homogeneous equation
(2.13)
(2.14)
P u :=
n
X
j=1
aj (x)
u
= f (x)
xj
dxj
= aj (x(t)),
dt
j = 1, . . . , n;
that is
dx
= V (x).
dt
17
u
= f (t, ).
t
This is exactly the special case (2.1) we have already solved. After one
has solved this, then one reverts to the original x coordinates.
Example: Solve 2ux + yuy uz = 0 with u(x, y, 0) = (x y)2 .
y = et
z = t + ,
y = ez .
The solution u(x, y, z) depends only on the integral curve passing through
the point (x, y, z), so it depends only on = x + 2z and = yez :
u(x, y, z) = h(x + 2z, yez )
for some function h which we now determine from the initial condition
2
Consequently,
u(x, y, z) = (x + 2z ye ) .
18
Lu := ux uy + cu = f (x, y).
i(k + c)uk = fk ,
fk
,
i(k ) + c
so
k, = 0, 1, 2, . . .
u(x, y) =
fk
ei(kx+y) .
i(k ) + c
4. A GLOBAL PROBLEM
19
i(k )uk = fk ,
k, = 0, 1, 2, . . .
R
If k = = 0 this implies f00 = 0, which is just T2 f = 0 dx (again).
Moreover, if = p/q is rational, then fk = 0 whenever k/ = p/q .
This gives infinitely many conditions on f . We will not pursue this case
further and consider only the case when is irrational. Then solving
(2.19) for uk and using them in the Fourier series for u we obtain
X ifk
(2.20)
u(x, y) =
ei(kx+y) .
k
It remains to consider the convergence of this series. Well use Lemma
2.39 to determine if u is smooth. It is clear that there will be trouble if
can be too-well approximated by rational numbers, since then the denominator (k/) will be small. This is the classical problem of small
divisors. Of course every real number can be closely approximated by
a rational number p/q . The issue is how large the denominator q must
be to get a good approximation.
Definition 2.3. An irrational number is a Liouville
number if for
k
p
every positive integer and any k > 0 then < for infinitely
q
q
many pairs of integers (p, q). Thus, is not a Liouville number if there
p
k
exist numbers and k so that > for all but a finite number
q
q
of integers p, q .
The results in the remainder of this section will not be used elsewhere in these
notes.
20
4. A GLOBAL PROBLEM
21
.
q2
1 + q2
1 + p2 + q 2
| fk |
c(s)(1 + k 2 + 2 )/2
c(s)
=
.
| k |
(1 + k 2 + 2 )s
(1 + k 2 + 2 )s(/2)
(1 +
p2
k
+ q 2 )/2
Using this with = 2j and k = 1, for each j pick one point (pj , qj ).
(2.22)
| pj qj | <
1
(1 + p2j + qj2 )j
2
We may assume that p2j + qj2 < p2j+1 + qj+1
to insure that each of these
lattice points selected is associated with only one index j . Define f by
setting
1
fpj qj =
2
(1 + pj + qj2 )j
22
for these lattice points (pj , qj ) while for all other lattice points (k, )
we set fk = 0. Then by Lemma 2.5 f is smooth. However from (2.19)
and (2.22)
| fpj qj |
| upj qj | =
>1
| pj qj |
so u is not smooth. In fact, u is not even in L2 (T2 ). Consequently if
is a Liouville number, then there is no smooth solution.
5. Appendix: Fourier series
Many problems in science and technology lead naturally lead one to
Fourier series. They are a critical tool in these notes.
a). Fourier series on S 1 . Say a function f (x) is periodic with period 2 . It is useful to think of these as functions on the unit circle, S 1 .
The simplest functions with this periodicity are eikx , k = 0, 1, 2 . . .
(or, equivalently, cos kx and sin kx). One tries to write f as a linear
combination of these functions
X
(2.23)
f (x)
a eix .
=
But how can you find the coefficients a ? What saves the day (and was
implicitly realized by Euler as well as Fourier) is to introduce the inner
product
Z
h, i =
(x)(x) dx,
In this inner product eikx and eix are orthogonal for integers k 6= .
As in Rn we also write the norm
Z
1/2
1/2
2
(2.24)
kk = h, i =
| (x) | dx
X
eix
(2.25)
f (s)
c .
2
=
Formally taking the inner product of both sides of this with eikx / 2
we obtain the classical formula for the Fourier coefficients
Z
eikx
eikx
(2.26)
ck = hf, i =
f (x) dx.
2
2
23
Understanding the convergence of the Fourier series (2.25) is fundamental. This convergence clearly depends on the decay of the Fourier
coefficients c . First we discuss convergence in the norm (2.24).
Let TN be the (finite dimensional) space of trigonometric polynomials
whose degree is at most N , that is, these functions have the form
P
P
eikx
ikx
tN (x) =
. Also let PN (f ) :=
| k |N ak e
| k |N ck 2 TN be
the terms in (2.25) with | k | N . By (2.26), note that f PN (f )
is orthogonal to TN because if | | N then hf PN (f ), eix i = 0.
Thus, we have written
f = PN (f ) + [f PN (f )]
k | ck
|2
(2.29)
kf PN (f )k kf tN k 2 ,
Remark 2.7. While this reasoning used that f C(S ), it is straightforward to see that the results hold only assuming f is piecewise continuous (or even square integrable). For this we use that in the norm
24
f (s)
eix
ic .
2
=
Consequently,
| ck |
2
max| f (x) |.
|k|
2
| ck |
max| D j f (x) |.
| k |j
Thus, the smoother f is, the faster its Fourier coefficients decay. In
particular, if f C 2 (S 1 ) (so f , f , and f are periodic), then | ck |
const /k 2 so the series | ck | converges and hence the Fourier series (2.25)
converges uniformly to f .
By being more careful, we can prove that the Fourier series converges
uniformly if f C 1 (S 1 ); in fact, all we will really require is that f is
square integrable. For this we use Bessels inequality (2.28) applied to
f:
X
(2.32)
| kck |2 kf k2 .
k
25
.
(1 + | k | )| ck |
1 + | k |2
| k |N
| k |N
The second series converges by (2.32) (in fact, it converges to kf k2 +
P
1/2
kf k2
), while the first by comparison to
1/| k |2 .
Exercises:
1. Let ck be the Fourier coefficients of f C(S 1 ). Show that if f and
all of its derivatives exist and are continuous, then for any integer
s 0 there is a constant M(s) so that | ck | M(s)/(1 + | k |2 )s/2 .
2. Conversely, if for any integer s 0 there is a constant M(s) so
that | ck | M(s)/(1 + | k |2 )s/2 , show that f C (S 1 ).
1
26
then, formally,
(2.36)
( + 1)u(x) =
X
k
(1 + | k |2 )uk eikx .
Using this and the divergence theorem we observe that for a real function u
Z
Z
X
2
2
2
u uu] dx =
| u | + | u | dx =
(1 + | k |2 )| uk |2 .
Tn
Tn
and
Z
Tn
2
| u | + 2| u |2 + | u |2 dx =
Tn
X
u(1 )2 u] dx =
(1+| k |2 )2 | uk |2
k
(for complex functions u one just adds a few complex conjugate signs).
Using this as motivation, define the Sobolev spaces H s (Tn ) to be the
space of functions with finite norm
X
kk2H 1 (Tn ) :=
(1 + | k |2 )s | k |2 < .
k
Of course H (T ) = L (T ).
One thinks of H s (Tn ) as the space of functions on Tn whose derivatives
up to order s are square integrable.
P To see this, letr r = (r1 , . . .r,1 rn ) be
any multi-index of integers with
rj = r and let D = (/x1 ) (/xn )rn
be a partial derivative of order r s. Then, using
X
k eikx
(2.37)
(x) =
k
we have
D r (x) =
X
k
kD r k2L2
X
k
| k |2r | k |2 .
27
(1
+
|
k
|
)
|
|
k
(1 + | k |2 )s
k
k
1/2
X
1
kkH 2 (Tn ) .
=
(1 + | k |2 )s
k
P
The series
1/(1 + | k |2 )s converges for all s > n/2. One way to see
this is by comparison with an integral using polar coordinates
Z
Z n1
dx
r
dr
n1
=
Area
(S
)
.
2 s
(1 + r 2 )s
Rn (1 + | x | )
0
This integral converges if 2s (n 1) > 1, that is, if s > n/2. Thus,
if s > n/2, there is a constant c so that if H s (Tn ) then
kkC 0 (Tn ) ckkH s (Tn ) .
If
s > j + n/2
then
28
30
CHAPTER 3
(3.3)
u(P ) + u(R) = u(Q) + u(S).
This is clear since u(P ) = F (a) + G(p), u(Q) = F (a) + G(q), u(R) =
F (b) + G(q), and u(S) = F (b) + G(p).
Light and sound are but two of the phenomena for which the classical
wave equation is a reasonable model. This study is one of the real success stories in mathematics and physics. It has led to the development
of many valuable techniques.
(3.4)
Upon studying the motion of a vibrating string one is led to the simple
differential equation
utt = c2 uxx ,
(3.6)
u = u .
b) Use this to solve (3.6) with the initial conditions
The term F (x ct) represents a wave traveling to the right with velocity c. We saw this in the previous Section a) when we discussed the
transport equation. The sketches there substantiate the statement that
c is the velocity of propagation of the wave. Similarly, G(x + ct) represents a wave traveling to the left with velocity c, so the general solution
is composed of waves traveling in both directions. The two families of
straight lines x ct =const, and x + ct =const are the characteristics
of the wave equation (3.1).
29
u = 0.
(3.2)
u(x, 0) = f (x)
ut (x, 0) = g(x).
(3.1)
initial position
initial velocity
(x,t)
x
x-ct
x+ct
Similarly, the initial data at a point (x0 , 0) can
only affect the solution u(x, t) for points in the
triangular region | x x0 | ct. This region is
called the domain of influence of the point (x0 , 0)
(x0 ,0)
31
32
u(x, 0) = f (x),
Special Case 2. A clever observation helps to solve the related problem for a semi-infinite string:
(3.8)
u(x, 0) = f (x), ut (x, 0) = g(x) for x > 0,
while u(0, t) = 0 for t > 0.
The observation is that for the infinite string < x < , if the
initial position u(x, 0) = f (x) and velocity ut (x, 0) = g(x) are odd
functions, then so is the solution u(x, t) (proof?). Thus, to solve (3.8)
we simply extend f (x) and g(x) to all of R as odd functions fodd (x)
and godd (x) and then use the dAlembert formula (3.5).
Exercise: Carry this out explicitly for the special case where (3.8)
holds with g(x) = 0. In particular, show that for x > 0 and t > 0
(
1
[f (x + ct) + f (x ct)] for x > ct
u(x, t) = 12
[f (ct + x) f (ct x)] for x < ct.
2
The boundary condition at x = 0 serves as a reflection. One can
see this clearly from a sketch, say with the specific function f (x) =
(x 2)(3 x)) for 2 x 3 and f (x) = 0 for both 0 x 2 and
x > 3.
can now be solved by simply adding the solutions from the two special
cases (3.7) (3.8) just treated.
u(x, 0) = f (x),
c). Finite string: 0 < x < L. In the case of a finite string, such
as a violin string, one must evidently also say something about the
motion of the end points x = 0 and x = L. One typical situation is
where we specify the position of these boundary points:
(3.9)
Thus, if the ends are tied down we would let f (t) = g(t) = 0. The
equations (3.9) are called boundary conditions. As an alternate, one
can impose other similar boundary conditions. Thus, if the right end
is allowed to move freely and the left end is fixed ((t) = (t) = 0),
then the above boundary conditions become
u
(3.10)
u(0, t) = 0
(L, t) = 0,
x
The condition at x = L asserts the slope is zero there (that the slope at
a free end is zero follows from physical considerations not given here).
There is no simple closed form solution of the mixed initial-boundary
value problem (3.1),(3.4), (3.9), even in the case f (t) = g(t) = 0. The
standard procedure one uses is separation of variables (see section c)
below). The solution is found as a Fourier series.
d). Conservation of Energy. For both physical and mathematical reasons, it is important to consider the energy in a vibrating string.
Here we work with an infinite string.
Z
(3.11)
E(t) = 12
(u2t + c2 u2x ) dx
u2t
The term
is for the kinetic energy and c2 u2x the potential energy.
(Here we have assumed the mass density is 1; otherwise E(t) should be
multiplied by that constant.) For this integral to converge, we need to
assume that ut and ux decay fast enough at . From the dAlembert
formula (3.5), this follows if the initial conditions decay at infinity.
We prove energy is conserved by showing that dE/dt = 0. This is a
straightforward computation involving one integration by parts in
33
which the boundary terms dont appear because of the decay of the
solution at infinity.
Z
Z
dE
2
(3.12)
ut (utt c2 uxx ) dx = 0,
(ut utt + c ux uxt ) dx =
=
dt
where in the last step we used the fact that u is a solution of the wave
equation.
Exercises
1. For a finite string 0 < x < L with zero boundary conditions:
u(0, t) = u(L, t) = 0, define the energy as
Z L
(3.13)
E(t) = 21
(u2t + c2 u2x ) dx.
0
and
initial position
u(x, y, 0) = f (x, y)
(3.17)
initial velocity
34
a). Formulas for the solution in R2 and R3 . There are standard formulas for the solution of the initial value problem (the term
Cauchy problem is often called).
Technical Observation Let x = (x1 , . . . , xn ) Rn . Say we want
to solve
(3.18)
utt = u,
vtt = v,
wtt = w,
and
(3.20)
Then vt also satisfies the wave equation but with initial conditions
vt (x, 0) = f (x) and vtt = 0. Thus the solution of (3.18) is u(x, t) =
vt (x, t)+w(x, t). Since both (3.19) and (3.20) have zero initial position,
one can find u(x, t) after solving only problems like (3.20). This is
utilized to obtain the following two formulas.
For the two (space) dimensional wave equation it is
(3.21)
ZZ
ZZ
1
f (, )
g(, )
1
u(x, y, t) =
d d +
d d,
2
2
2
2c t
2c
c t r
c2 t2 r 2
rct
rct
where r = (x ) + (y ) .
r=ct
Bt = curl E,
div B = 0,
div E = 0.
35
div E(x, 0) = 0, then div B(x, t) = 0 and div E(x, t) = 0 for all
t > 0.
n
where the coefficients ajk are constants and (without loss of generality why?) akj = ajk . If the matrix A = (ajk ) is positive
definite, show there is a change of variable x = Sy , where S is
an n n invertible matrix, so that in these new coordinates the
equation becomes the standard wave equation
n
2u X 2u
.
=
t2
2 y
=1
b). Domain of dependence and finite signal speed. As before, it is instructive to examine intersection of the domain of dependence with the plane t = 0, in other words, to
determine the points x for which the initial data
can influence the signal at a later time. In the two
dimensional case (3.21), the intersection of the domain of dependence of the solution at (x0 , y0 , t0 )
t
with the plane t = 0 is the entire disc r ct0 ,
(x,y,t)
while in the three dimensional case (3.22), the domain of dependence is only the sphere r = ct0 ,
not the solid ball r ct0 . Physically, this is intery
ct
preted to mean that two dimensional waves travel
with a maximum speed c, but may move slower, x
while three dimensional waves always propagate
with the exact speed c.
This difference in observed in daily life. If one drops a pebble into
a calm pond, the waves (ripples) move outward from the center but
ripples persist even after the initial wave has passed. On the other
hand, an analogous light wave, such as a flash of light, moves outward
as a sharply defined signal and does not persist after the initial wave
has passed. Consequently, it is quite easy to transmit high fidelity
waves in three dimensions but not in two. Imagine the problems
in attempting to communicate using something like Morse code with
waves on the surface of a pond.
For the two space variable wave equation, the characteristics are the
surfaces of all light cones (x )2 + (y )2 = c2 t2 . In three space
dimensions, the characteristics are the three dimensional light cones.
They are the hypersurfaces in space-time with (x )2 + (y )2 +
(z )2 = c2 t2 .
36
utt = c2 u
where
u = ux1 x1 + + uxn xn ,
u(x, 0) = f (x),
ut (x, 0) = g(x)
37
1
2
D(t)
D(t2 )
(u2t + c2 | u |2) dx
1
2
D(t1 )
c(T t)
Z
S(r)
u2t + c2 | u |2 dr
D(t)
ut u dx,
D(t)
D(t)
Next, we note that utt = c u so the first integral is zero. For the
second term we use the standard inequality 2ab a2 + b2 for any real
a, b to obtain the estimate
dr,
1
2
where S(r) is the n1 sphere (in the plane) with radius r and centered
at (X, t). Hence
Z
Z
dE
ut utt + c2 u ut dx 2c
=
u2t + c2 | u |2 dc(T t) .
dt
D(t)
S(c(T t)))
E(t) =
38
39
c). Mixed Initial-Boundary Value Problems. The above formulas (3.21), (3.22) were for waves in all of space. In the case of a
vibrating membrane , we must also impose boundary values on ,
the boundary of . Similarly, in the case of light or sound waves outside of , we put boundary conditions on both and at infinity
(this is sometimes referred to as an exterior problem, while a vibrating
membrane is an interior problem. Just as for the vibrating string (...
), two typical boundary conditions are
u(x, t) = f (x, t) for x
(Dirichlet conditions)
(3.25) u
(x, t) = g(x, t) for x
(Neumann conditions),
where / means the directional derivative in the outer normal direction to . Of course this presumes that the boundary is smooth
enough to have an outer normal direction. One also has situations
where one of these conditions holds on part of the boundary and the
other on another part. The vibrating string (3.10) is an example.
We now restrict our attention to waves in a bounded region , such
as a vibrating membrane, and use the method of separation of variables to solve the wave equation with homogeneous Dirichlet boundary
conditions:
(3.26)
utt = u
(3.26a)
u(x, t) = 0
(3.26b)
u(x, 0) = f (x),
for
x ,
ut (x, 0) = g(x),
Since the left side depends only on x while the right depends only on
t, they must both be equal to a constant, say . Thus we obtain the
two equations
(3.28)
T T = 0
W = W.
40
over :
W 2 (x) dx =
W (x)W (x) dx =
| W |2 dx,
W = W
in
w = 0 on
| W |2 dx
| W |2 dx
Thus is an eigenvalue of the operator, with corresponding eigenfunction W (x). For a membrane , these eigenvalues are essentially
the squares of the various frequencies with which the membrane can
vibrate and the eigenfunctions are the normal modes. It turns out that
only a discrete sequence of eigenvalues 0 < 1 2 are possible
with j as j . Write the corresponding
eigenfunctions
as
p
p
j . From (3.28) the functions Tj (t) = aj cos j t + bj sin j t so the
standing wave solutions (3.27) are
p
p
(3.31)
uj (x, t) = aj cos j t + bj sin j t j (x),
W (0) = 0,
W () = 0.
ut x, 0 = 0,
u(0, t) = u(, t) = 0.
(3.33)
1 = min Z
,
| |2 dx
where the minimum is taken over all C 1 functions that satisfy the
Dirichlet boundary condition = 0 on . Assuming there is a function C 2 ()) C 1 () that minimizes (3.30), we will show that it is
an eigenfunction with lowest eigenvalue 1 . To see this say such a
minimizes the functional
Z
| v |2 dx
J(v) = Z
,
| v |2 dx
42
6. SMOOTHNESS OF SOLUTIONS
43
Using the variational characterization(3.33), it is easy to prove a physically intuitive fact about vibrating membranes: larger membranes have
a lower fundamental frequency. To prove this, say + are bounded
domains with corresponding lowest eigenvalues 1 () and 1 (+ ).
Both of these eigenvalues are minima of the functional (3.33), the only
difference being the class of functions for which the minimum is taken.
Now every admissible function for the smaller domain is zero on
and hence can be extended to the larger domain by setting it to be zero
outside . It is now also an admissible function for the larger domain
+ . Therefore, for the larger domain the class of admissible functions
for J(v) is larger than for the smaller domain . Hence its minimum
1 (+ ) is no larger than ().
Using similar reasoning, one can prove a number of related facts, and
also get explicit estimates for eigenvalues. For instance, if we place
R2 in a rectangle + , since using Fourier series we can compute
the eigenvalues for a rectangle, we get a lower bound for ().
6. Smoothness of solutions
From the formula u(x, t) = F (x ct) + G(x + ct) for the solution of the
one dimensional wave equation, since formally F and G can be any
functions, it is clear that a solution of the wave equation need not be
smooth (this is in contrast to the solutions of the Laplace equation, as
we shall see later). In fact, in higher dimensions, even if the initial data
(3.16) are smooth, the solution need not even be continuous. This can
be seen intuitively for three space variables by choosing initial conditions on a sphere so that light rays are focuses at the origin at a later
time. This is commonly done with a lens. To see this with formulas,
notice that for any smooth f C (R) the function
u(x, y, z, t) =
f (r + ct)
,
r
where r 2 = x2 + y 2 + z 2 ,
2
44
The approach is analogous to Lagranges method of variation of parameters, which gives a formula for the solution of an inhomogeneous
equation such as u +u = F (t) in terms of solutions of the homogeneous
equation. The method is called Duhamels principle.
We illustrate it for the wave equation, seeking a solution of (3.35) with
initial conditions
u(x, 0) = 0 ut (x, 0) = 0.
Since we are solving a differential equation, it is plausible to find a
solution as an integral in the form
Z t
v(x, t; s) ds
(3.36)
u(x, t) =
0
The similar formula for u is obvious. Substituting these into the wave
equation (3.35) we want
Z t
Lv(x, t; s) ds + vt (x, t; s)s=t .
F (x, t) = Lu(x, t) =
0
This is evidently satisfied if Lv = 0 and vt (x, t; s)s=t = F (x, t) along
with v(x, t; t) = 0 for all t 0.
Because the coefficients in the wave equation do not depend on t, our
results can be simplified a bit by writing v(x, t; s) = w(x, t s; s) so
for each fixed s, the function w(x, t; s) satisfies
(3.37) wtt = c2 w
45
We can now find w by using our earlier formulas. For instance, in three
space variables, from (3.22)
ZZ
1
F (, s) dA ,
w(x, t; s) =
4c2 t
kxk=ct
1
4c
kxk=c(ts)
ZZ
kxk=c(ts)
F (, t k xk/c)
dA ds.
k xk
x R,
v(r, 0) = (r),
vt (r, 0) = (r).
46
48
ut = ku,
ut = uxx
with
u(x, 0) = f (x).
Assuming a mild growth condition on f , say it is bounded and continuous, the solution is
Z
(xs)2
1
(4.3)
u(x, t) =
f (s)e 4t ds.
4t
Before going further it is useful to make some observations based on this
formula. First, it implies that if the initial temperature is non-negative
but not identically zero, then the solution is positive everywhere, even
for very small t. Thus, in contrast to the solution of the wave equation,
heat conduction has an infinite signal speed. We also observe that even
if f is, say, only piecewise continuous, the solution is smooth in both
x and t for all t > 0. In fact, it has a power series in x that converges
47
To derive (4.3) for the moment we work formally and assume all integrals make sense. First take the Fourier transform of ut = uxx with
respect to the space variable x,
Z
u(x, t)eix dx.
u(, t) :=
Then from (4.27) in the Appendix to this chapter u(, t) satisfies the
ordinary differential equation
ut = | |2 u
with
u(, 0) = f()
in which appears only as a parameter. Its solution is
2
u(, t) = e| | t f(t).
Z
| xy |2
1
=
f (y)e 4t dy.
4t
This derivation was purely formal. Since the resulting formula may well
hold under more general conditions than this derivation admits, instead
of checking each step we verify directly that it solves the heat equation
and satisfies the initial condition. By differentiating under the integral
we immediately verify that it satisfies the heat equation ut = uxx for
all t > 0. Moreover u is a smooth function of x and t for all t > 0.
It remains to show that limt0 u(x, t) = f (x). This is a special case of
the next lemma.
Let C(R) have the properties
(1)
R (x) 0,
(2) R (x) dx = 1,
R
(3) For any > 0, lim0 | y | (y) dy = 0.
Let
(4.4)
f (x) :=
f (y)(x y) dy.
2. SOLUTION FOR Rn
49
| z |<
Thus,
| f(x) f (x) | < + 2M
| z |
ut u = F (x, t)
By Property 3) the last integral can be made arbitrarily small by choosing sufficiently small. Since the right hand side is independent of x
(as long as x K ), the convergence is uniform.
(4.5)
50
u(x, t) =
with u(x, 0) = 0.
v(x, t; s) ds.
with
4. MAXIMUM PRINCIPLE
51
initial temperature
u(x, 0) = f (x)
(4.9)
boundary temperature
u(x, t) = g(x, t)
for x
for x .
u(x, t)
= g(x, t),
N
for x
52
and
(4.12)
w(x, 0) 0 for
while
w(x, t) 0 for x ,
0tT
v(x, 0) = g(x)
(4.15)
u(x, t) = (x, t)
for x
v(x, t) = (x, t)
for x , t > 0.
4. Maximum Principle
(4.11)
wt w 0
for x T ,
for all 0 t T,
4. MAXIMUM PRINCIPLE
53
(4.17)
t 0.
and
u(x, t) = (x, t)
(x , t > 0).
and
| | <
for
x , t 0.
54
theorem for the initial value problem (4.5) for the heat equation in all of
Rn . We prove uniqueness for this in the class of bounded solutions (one
2
can weaken this to allow u(x, t) consteconst | x | , see [PW], p. 181,
but there are examples of non-uniqueness if one allows faster growth).
Say u(x, t) satisfies the heat equation ut = u in all of Rn with
u(x, 0) = 0 and | u(x, t) | M . Inside the disk {| x | < a} consider
the comparison function v(x, t; a) := M(| x |2 + 2nt)/a2 . Then v also
satisfies the heat equation with
v(x, 0; a) 0 while v(x, t; a) M u(x, t) for | x | = a, t > 0.
Thus by the maximum principle u(x, t) v(x, t; a) for | x | a. Fixing
(x, t) but letting a we conclude that u(x, t) 0. Replacing u
by u we then get u(x, t) = 0 for all t 0.
Exercises
1. Let u(x, t) be a bounded solution of the heat equation ut = uxx
with initial temperature u(x, 0) = f (x). If f (x) is an odd function
of x R, show that the solution u(x, t) is also an odd function of
x.
2. Semi-infinite interval Solve the heat equation on a half-line:
0 < x < with u(x, 0) = f (x) for x 0 and the following
conditions:
a) u(x, 0) = f (x) for x 0 and u(0, t) = 0 for t 0. [Suggestion: Extend f (x) cleverly to x < 0.]
b) u(x, 0) = 0 and u(0, t) = g(t). [Suggestion: Let v(x, t) =
u(x, t) g(t).]
c) u(x, 0) = f (x) for x 0 and u(0, t) = g(t) for t 0.
5. Prove Corollary 7.
b). Symmetry of solutions. Uniqueness is often the easiest approach to show that a solution possesses some symmetry. One example makes the ideas transparent. Let R2 be the rectangle
{ | x | < 1, 0 < y < 1 } and let : be the reflection across
the y -axis. Assume the initial and boundary temperatures are invariant under , so they are even functions of x. We claim the solution is also invariant under . This is obvious since both u(x, t) and
v(x, t) := u((x), t) are solutions of the heat equation with the same
initial and boundary values.
c). Uniqueness in Rn . If is unbounded, such as an infinite rod
{ < x < }, then the simple example u(x, t) := 2t + x2 which
satisfies the heat equation but whose maximum does not occur at t = 0
shows that the maximum principle fails unless it is modified. However,
we can still use it as a tool. To illustrate, well prove a uniqueness
56
(4.20)
u(x, t) =
(4.21)
=
f (y)G(x, y)dy,
XZ
where G(x, y, t) =
vk (y)vk (x)ek t .
Here G(x, y, t) is called Greens function for the problem. Because the
eigenvalues, k are all positive, it is clear from (4.19) that u(x, t) 0
as t tends to infinity. This should agree with your physical intuition.
The lowest eigenvalue, 1 , determines the decay rate.
Exercise: Repeat this using homogeneous Neumann boundary conditions u/N = 0 on the boundary. What can you say about limt u(x, t)?
b). Another approach. Using techniques similar to the energy
methods we used for the wave equation, we can also obtain information
about solutions of the heat equation. These are reasonable exercises.
Exercises
1. Let u(x, t) be a solution of the heat equation ut = u in with
u = 0 on . Define
Z
H(t) := 12
u2 (x, t) dx.
To derive the standard formula for the solution of the heat equation of
an infinite rod, we used the Fourier transform. Here is a brief summary
of basic facts about the Fourier transform. If u L1 (Rn ), its Fourier
transform u() is defined as
Z
u(x)eix dx,
(4.22)
u() :=
Rn
It is evident that | u
() | kukL1 (Rn ) .
a). A special integral.
1
xAx+bx
e
dx =
(4.23)
I :=
ebA b/4 ,
det
A
n
R
2
2
e d = .
e d =
iq
Combined with the above formula for I this gives the desired formula.
57
| x |2 /2
Rn
| x |2 /2
Thus e
Now
easy computation its Fourier transform is (y) = n (y/).
(4.26)
Z Z
Z
u(y)eiy dy ()eix d
u() ()eix d =
Rn
Rn
Z
Z
u(y)
()ei(yx) d dy
=
n
Rn
Z
ZR
dt.
u(x + t)(t)
u(y) (y x) dy =
=
Rn
Rn
Rn
| x |2 /2
Choosing (x) = e
and using (4.24) gives the desired Fourier
inversion formula, at least for bounded functions u L1 (R).
c). Fourier transform of the derivative. One reason the Fourier
transform is so useful when discussing linear differential equations with
constant coefficients is that the Fourier transform changes differentiation into multiplication by a polynomial. This is easily seen by integrating by parts
Z
j u(x)eix dx = ij u().
(4.27)
c
j u() =
[
In particular, (u)()
= | |2 u(). so for any integer k 0
(1 + | |2 )k u() = [(1 \
)k u]().
58
60
2. Poisson Equation in Rn
CHAPTER 5
u = 0.
u(x )
.
| x |n2
= 0 ,
61
62
To prove this, let d be the element of area on the unit sphere; then
on B(x0 , r) we have dA = r n1d ,
(5.6)
0=
Z
u
u(x + r)
dA = r n1
d
r
B(x,r) N
| |=1
Z
d
u(x + r) d .
= r n1
dr
| |=1
u(y) dy =
B(x,r)
(5.7)
which is (5.5).
The solid mean value property is
Z
1
u(x) =
u(y) dy.
Vol (B(x, r)) B(x,r)
It follows from the mean value property for spheres by simply multiplying both sides by Area (B(x, r)) and integrating with respect to
r.
The maximum principle is an easy consequence. It asserts that if u
with u = on
Exercises:
1. Show that this second proof also works with Neumann boundary
conditions u/N = 0, except that with these boundary conditions we can only conclude that u constant. Indeed, if u is any
solution, then so is u + const.
2. If in a bounded domain say u = 0 with u = f on the boundary
while v = 0 with v = g on the boundary. If f < g what can you
conclude? Proof?
3. If u satisfies u 0, show that the average of u on any sphere
is at least its value at the center of the sphere. Use this to conclude
that if u 0 on the boundary of a bounded domain , then u 0
throughout .
4. In a domain Rn let u(x) be a solution of u + a(x)u = 0,
where a(x) > 0.
a) Show there is no point where u has a positive local maxima (or
negative minima).
b) In a bounded domain, show that there is at most one solution
of the Dirichlet problem
+ a(x)u = F (x) in with u = on .
[Give two different proofs, one using part a), the other using
Energy.]
5. In a domain Rn let the vector u(x) be a solution of the system
of equations u + A(x)u = 0 with u = 0 on the boundary.
Here A(x) is a symmetric matrix and u means apply to each
component of u; a useful special case is the system of ordinary
differential equations u + A(x)u = 0. Assume A(x) is a positive
definite matrix, show that u 0. Also, give an example showing
that if one drops the assumption that A(x) is positive definite, then
there may be non-trivial solutions. Suggestion: As just above,
there are two distinct approaches, both useful:
i). Use energy methods directly.
63
ii). Let (x) = | u(x) |2 and apply the scalar maximum principle to .
2
n
X
i,j=1
aij
2u
+ c(x)u = 0,
xi xj
0 u(x1 )
Rn
u(x0 ).
(R | x1 |)n
To prove this, since the ball B(x1 , R| x1 |) B(x0 , R) we use the solid
mean value property in B(x1 , R | x1 |), the assumption that u 0,
and the mean value property a second time to find
Z
1
u(x) dx
u(x1 ) =
Vol (B(x1 , R | x1 |) B(x1 , R| x1 |)
Z
1
u(x) dx
Vol (B(x1 , R | x1 |) B(x0 ,R)
=
64
u = 0 in with u = f on .
The function
P (x, y) =
is called the Poisson kernel.
R 2 | x |2
nn R| x y |n
There are several ways to derive (5.10). If n = 2 one can use separation
of variables in polar coordinates. Other techniques are needed in higher
dimensions. The details are carried out in all standard texts.
The mean value property is the special case of (5.10) where x = 0.
One easy, yet important, consequence of the Poisson formula is that if
a function u is harmonic inside a domain , then it is smooth (C )
there. To prove this near a point x consider a small ball B(x0 , R)
containing x and use (5.10) to obtain a formula for u in terms of its
values on the boundary of the ball:
Z
R 2 | x |2
u(y)
u(x) =
dA(y).
nn R
|
x
y |n
B(x0 ,R)
65
66
having a Fourier series (2.35), then from (2.36), matching the coefficients we find that
fk
uk =
1 + | k |2
so
X
fk
(5.12)
u(x) =
eikx .
2
1
+
|
k
|
k
means
lim g(s)/h(s) = 0.
ss0
For example x2 = o(x) as x 0. In this notation, the precise assumption needed on u(x) for a removable singularity is
u(x) = o((x)) as x 0.
Exercises:
1. Use separation of variables in polar coordinates to obtain the Poisson formula for the unit disk in R2 .
2. Use separation of variables in polar coordinates to solve the Dirichlet problem for the annulus 0 < a2 < x2 + y 2 < 1 in R2 .
3. Let uk be a sequence of harmonic functions that converge uniformly
to some function u(x) in a domain . Show that u is also harmonic.
4. [Harnack inequality] Let u(x) 0 be harmonic in the ball
B(0, R). Use the Poisson formula to show that
Rn2
R |x|
R + |x|
u(0) u(x) Rn2
u(0).
(R + | x |)n1
(R | x |)n1
(5.11)
u + u = f
on Tn
We summarize this.
v + v = (u + 2 u + u) + u = G,
G = f 2 u u.
67
68
polynomials:
dim H = dim P dim P2 =
(n + 2 2)(n + 3)!
n+1
n+3
=
.
!(n 2)!
(5.15)
69
Application. Atoms are roughly spherically symmetric. The maximum number of electrons in the k th atomic subshell is related to the
dimension of the eigenspace corresponding to the k th eigenvalue. The
Pauli exclusion principle asserts that no two electrons can be in the
same state. But electrons can have spins 1/2, There are 2k + 1 electrons with spin 12 , so 2(2k + 1) in all. Thus the subshells contain at
most 2, 6, 10, 14, . . . electrons.
7. Dirichlets principle and existence of a solution
a). History. To solve the Dirichlet problem (5.9), Dirichlet proposed to find the function u that minimizes the Dirichlet integral
Z
(5.16)
J() := | |2 dx
Since h can be any piecewise smooth function that is zero on the boundary, this implies that u = 0, as desired. [Proof: if not, say u > 0
somewhere, then u > 0 on a small ball. Pick a function h that
is
R positive on this ball and zero elsewhere, giving the contradiction
u h dx > 0.]
Riemann adopted this reasoning in his proof of what we now call the
Riemann mapping theorem. Weierstrass pointed out that although
J(u) is bounded below and hence has an infimum, it is not evident that
that there is some function u satisfying the boundary conditions for
which J(u) has that minimal value. To make his argument convincing,
he gave the example
Z 1
x2 (x)2 dx
with (1) = 1.
J() =
1
1 for 1 x 1/k,
k (x) = kx for 1/k x 1/k,
1 for 1/k x 1.
70
2
Then J(k ) = 3k
0, so inf J() = 0. But if J(u) = 0, then
u = const and cant satisfy the boundary conditions.
Since Riemanns application of Dirichlets principle was important,
many people worked on understanding the issues. Using other methods
Poincare gave a rather general proof that one could solve the Dirichlet
problem (5.9) while around 1900 Hilbert showed that under reasonable
conditions, Dirichlets principle is indeed valid.
,
b). A modified problem. In subsequent years the tools developed to understand the issues have led to a considerable simplification.
First, instead of solving (5.9) solve the related inhomogeneous equation
(5.18)
u = F in
with
u = 0 on .
for all h that vanish on the boundary. As before, assuming this function u has two continuous derivatives, an integration by parts shows
that u = F , as desired. It is not difficult to show that Q is bounded
below, but even knowing this we still dont know that Q achieves its
minimum. Instead of perusing this, we take a slightly different approach.
For a bounded open set , use the space Cc1 () of functions with
compact support in (the support of a function is the closure of the
set where the function is not zero). Thus, the functions in Cc1 () are
zero near the boundary of . For Cc1 () define the norm
Z
kk2H 1 () = | |2 dx.
0
Because of the Poincare inequality (3.34), this is a norm, not a seminorm. Define the Sobolev space H01 () as the completion of Cc1 () in
71
72
H01 ().
R
Letting v = w we conclude that | w |2 =0 so, using the Poincare
inequality (3.34), w = 0. Consequently, if we have a weak solution and
if we believe there is a classical solution, then the only possibility is
that the weak solution is also the desired classical solution.
Our strategy is to break the proof of the existence of a solution into
two parts:
Existence: Prove there is a weak solution.
Regularity: Prove that this weak solution is a classical solution
if f is smooth enough.
Consequently, if we let z = (x1 )e1 + + (en )en , then (x) = hx, zi.
Geometrically one can interpret z as a vector orthogonal to the kernel
of .
Essentially the same proof works in any separable Hilbert space. Pick
a (countable) orthonormal basis and write x H in this basis. Then,
as above, we are led to let
z = (x1 )e1 + + (en )en + .
just as desired.
Note that this proof works for any bounded open set , no matter
how wild its boundary. For instance, if is the punctured sphere
0 < kxk < 1 in R3 and try to solve u = 4 there with u = 0 on the
boundary, the unique solution in polar coordinates is u = (1 | r |2 )
which does not satisfy the boundary condition we attempted to impose,
u(0) = 0. That jump discontinuity is a removable singularity. The
existence theorem is smart enough to ignore bad points we may have
on the boundary of .
Exercises:
73
H01 ()
Z
with u = 0 for x .
to be a weak solution of Lu = f if
Z
f v dx
[a(x)u v + c(x)uv] dx =
for all v H01 (). Prove that there exists exactly one weak soDefine and use a Hilbert space that uses
Rlution. [Suggestion:
[a(x)
+
c]
dx
as
its inner product. Show that the norm on
u
Lu :=
aij (x)
+ c(x)u = F (x)
xi
xj
i,j=1
where 0 c(x) is continuous and F L2 (). Show there is
a unique weak solution u H01 () of Lu = F L2 (). [The first
step is to define a weak solution].
74
CHAPTER 6
The Rest
In the last part of the course I outlined the several topics, mainly following various parts from my old notes
Lecture Notes on Applications of Partial Differential Equations to Some
Problems in Differential Geometry, available at
http://www.math.upenn.edu/kazdan/japan/japan.pdf
In addition, there is a bit of overlap with my expository article Solving
Equations available at
http://www.math.upenn.edu/kazdan/solving/solvingL11pt.pdf
Topics
Defined both the Holder spaces C k+ , 0 1 and Sobolev
spaces H p,k and illustrated how to use them in various regularity assertions for solutions of some linear and nonlinear elliptic
partial differential equations.
Defined ellipticity for nonlinear equations, giving several examples including a Monge-Amp`ere equation.
Discussed issues concerning qualitative properties and existence for the minimal surface equation, equations of prescribed
mean and Gauss curvature (for surfaces) and some equations
for steady inviscid fluid flow.
Discussed techniques for proving that a partial differential equation has a solution. The techniques included:
a) iteration using contracting mappings,
b) direct methods in the calculus of variations,
c) continuity method
d) fixed point theorems (Schauder and Leray),
e) heat equation (R. Hamilton).
75