C ha pter 2
L inear First-Or der PDEs
The general first-order partial differential equation (PDE) for a two-variable function, denoted as
u=u(x; y), can be expressed in the form:
F (x; y; u; u
x
; u
y
) = 0:
Here, u
x
and u
y
repre sent the partial derivatives of u with respe ct to x and y, respectively. The
function F establishe s a functional relationship between the function u, its partial derivatives, a nd
the independent variables x and y. The general first-order partial differential equation f or a function
u=u(x; : : : ; x) o f n independent variables, denoted as x; : : : ; x, can be represented as:
F (x; u; ru)=0;
Here, ru is a vector denoted as ru= (@u; : : : ; @u), which comprises the partial derivatives of u
with respect to each independent variable x; : : : ; x.
Definition 2.1. A classical solution of the equation F (x; u; ru)=0, for x2R, ru= (@u; :::; @u),
is a smooth function u=u(x) defined on an open set R such that F (x; u(x); ru(x))=0, is an
identity for all x 2.
For example, it is possible to verify that functions of the form u =h(x
2
+ y
2
) for arbitrary smooth
functions h is the classical solution of the equation
yu
x
¡xu
y
=0:
For example, the function u = x
2
+ y
2
is a classical solution to the equation for all (x; y) 2R
2
, while
u = x
2
+ y
2
p
is a solution only on R
2
¡f(0; 0)g. The gr aph of these two solution shown below.
Observe that the graph of a classica l solution of a first-order PDE in two variable x; y is a smooth
surface:
1
2.1 Cl assification of first-order PDEs
In this book, we will exclusively focus on the study of first-order PDEs falling within the categories
of linear, semi-linear, and quasi-linear. The subject of fully nonline ar equations will be introduced
in a separate book dedicated to that topic.
Linear equations. The general form of a linear first-order PDE for a function u=u(x), x =
(x
1
; : : : ; x
n
) is given by:
X
j=1
n
v
j
(x) @
j
u(x) + v
0
(x) u = r(x);
for some (usually) continuous functions v(x) and r(x).
Semi-linear equations. A semi-linear equation is characterized by the general form:
X
j=1
n
v
j
(x) @
j
u(x) = r(x; u);
The difference between a linear and semi-linear equation is that a semi-linear equation can
be nonlinear with respect to u (and not with the partial deriva tives u
x
and u
y
).
Quasi-linear equations. A quasi-linear equation assumes the general form:
X
j=1
n
v
j
(x; u) @
j
u(x) = r(x; u);
The difference between a quasi-linear and semi-linear equation is that in the former case, the
coefficients of partial derivatives are function of u a s well.
Fully-nonlinear equations. A fully nonlinear equation is an equation where one or all of the
partial derivatives are nonlinear. For example, the equation:
ju
x
j² + ju
y
j² =1;
is a fully nonlinear first-order equatio n for u=u(x; y). We wont study this type of equation
in this book.
Exercise 2.1. Classify the following rst-order eq uations
a) u
x
+ u
y
= e
x
u
b) xu
x
+ yu
y
= e
u
c) u
x
+ (u
2
)
x
= 1 + u
d) u
x
u
y
+ uu
z
= 1
Exercise 2.2. Consider t he linear equation
u
x
+ u
y
= ¡u:
Verify that every fun ction of the forms u = f(y ¡ x) e
¡x
, u = f (y ¡x) e
¡y
satisfies the equation, where f is a
smooth arbitrary function.
Exercise 2.3. Consider t he following quas i- linear equation
u
x
+ uu
y
= 0:
a) Verify tha t every function of the implicit form u = f(y ¡ux) solves the equation.
b) What is t he exp licit solution if we know u satis es the con dition u = 2y + 1 along the y-axis ?
Exercise 2.4. Verify that the function u = 1 ¡ x
2
+ y
2
p
is a solution t o the fully nonlinear equation
ju
x
j
2
+ ju
y
j
2
= 1:
What is the domain of u if it considered as a c lassical solution? The solu tion satisfies th e auxiliary condition
u = 1 on the unit circle x
2
+ y
2
= 1. Note that u = x
2
+ y
2
p
¡1 solves the PDE and the auxiliary condition too.
2 Li near First-Order PDEs
2.2 Characteristic method and OD Es along cu rves
The characteristic method is a powerful technique for solving first-order partial differential equa-
tions, and it is especially useful for semi-linear and quasi-linear equations. By using this method,
one can derive the general solution to such equations. Furthermore, the characteristic method has
a geometric interpretation that can be illustrated through the Cauchy problem.
2.2.1 Introductory remark: ODE along a curve
In our study of ordinary differential equations (ODEs), we explore d equations of the form:
du
dx
= f (x; u):
Here, u = u(x) represents a single-variable function. Geometrically, we interpret the x-variable as the
x-axis in the standard direction. The solution to this equation consists of a one-parameter family of
functions u = u(x; c), where c 2R, such that for any x within the domain of u, the following relation
holds:
d
dx
u(x; c) = f(x; u(x; c)):
Now, let's shift our focus to a parametric curve γ(t) in the xy-plane. An ordinary differential
equation (first-order) along γ(t) takes the form:
du γ
dt
= f (t; u γ):
Here, uγ is defined at any t as (uγ)(t)= u(γ(t)). If we denote w as uγ, we arrive at the equation:
dw
dt
= f (t; w):
γ(t) = (x(t); y(t))
x
dw
dt
= f(t; w)
y
For instance, consider γ(t) given by γ = (t; t
2
), a parabola in the xy-plane along which the
differential equati on:
dw
dt
= w;
is defined. Suppose u at the point (1 ; 0) is 1, corresponding to γ(0). Then, we obtain w as:
w(t) = u(γ(t)) = e
t
:
2.2 Characteristic method and ODEs along curves 3
Remark 2.2. Solving differential equations along curves can sometimes result in non-valid solu-
tions. For instance , let's consider the circle C represented by the parametric curve:
γ(t) = (cos(t); sin(t));
for t in the interval [0; 2π), along with the initial value problem:
dw
dt
= w; w(0) = 1;
defined for w(t) = u(γ(t)). The solution to this equation is w(t) = e
t
for t 2[0; 2π). However, the
function u is not continuous on the circle because:
u(γ(0)) = u(1; 0) = 1;
and
u
lim
t!2π
γ(t)
= u(γ(0)) = 1 =/ lim
t!2π
u(γ(t))
Exercise 2.5. Consider the famil y of د fy = x
2
+ c; c 2Rg for parameter c 2R. Consider th e ord inary differential
eq uation
du
dx
= 0;
alo ng all curves i n this family. Show that u on the xy-plane can be descri bed by u = h(y ¡ x
2
) where h is an
ar bitrary smooth function.
2.2.2 A simple type of e quations
Let's begin with the following simple equation:
u
x
+ v(x; y) u
y
= 0; (2.1)
where u is a smooth two-variable function, u = u(x; y). We'll relate the independent variable y to
x through the equation:
dy
dx
= v(x; y);
and assume that the solution to this equation is expressed as y = Y (x; c), where c is a para meter
of the solution t o this ordinary differential equation (ODE). This family of curves is known as the
characteristic curves of the given partial differential equation (PDE). The reason is that along each
curve y = Y (x; c), the PDE reads as an ODE:
d
dx
u(x; Y (x; c)) = 0: (2.2)
Note that, by the chain rule, we have:
d
dx
u(x; Y (x; c)) = u
x
(x; y) +
dy
dx
u
y
(x; y) = u
x
(x; y) + v(x; y) u
y
(x; y):
4 Li near First-Order PDEs
Equation (2.2) can be simply solved for the constant function:
u(x; Y (x; c)) = C ;
where C is constant along the characteristic curve y = Y (x; c) for a fixed c. Therefore, C is a f u nction
of c, written as C = h(c), where h is an arbitrary function.
u = C
3
y = Y (x; c
2
)
y = Y (x; c
1
)
y
x
u = C
2
y = Y (x; c
3
)
u = C
1
du
dx
(x; Y (x; c)) = 0
Let's assume that the equation y = y(x; c) can be solved for c as c = g(x; y). Then, we can
express u(x; y) as
u(x; y) = h(g(x; y)); (2.3)
for an arbitrary smooth function h. Now, let's verify that the solution (2.3) satisfies equation (2.1):
(
u
x
= h
0
(g(x; y)) g
x
u
y
= h
0
(g(x; y)) g
y
:
This implies:
u
x
+ v(x; y) u
y
= h
0
(g(x; y))(g
x
+ v(x; y) g
y
)
Utilizing the relation c = g(x; y), we have
0 = g
x
dx + g
y
dy;
which leads t o:
g
x
+
dy
dx
g
y
= 0:
Substituting this into our previous equation:
(h(g(x; y)))
x
+ v(x; y)(h(g(x; y)))
y
= 0:
This confirms that the solution (2.3) satisfies equation (2.1).
Example 2.3. Consider the partial differential equation:
u
x
+ u
y
= 0:
To apply the characteristic method, we begin by finding the characteristic equation:
dy
dx
= 1:
This equation has the solution y = x + c, where c is a parameter. Now, let's consider the characteristic
family, denoted as fc = y ¡x; c 2Rg. Along each characteristic curve, u remains constant due to
the equation:
du
dx
= 0. Thus, we can express u as u = C along each line c = y ¡x, where C depends
on c. Consequently, we have:
u(x; y) = h(y ¡x):
2.2 Characteristic method and ODEs along curves 5
Here, h is an arbitrary smooth function. The characteristic curves in this case are straight lines
with a slope of 1.
x
c
0
y
(x; y)
c
1
c
¡1
u = C
1
= h(c
1
)u = C
0
= h(c
0
)
u = C
¡1
= h(c
¡1
)
x ¡ y
Remark 2.4. In the example we solved earlier, we obtained the solution u in terms of an arbitrary
function h(y ¡x). Consequently, any function of the form u =sin(y ¡x), u = e
¡(y ¡x)
2
, u = (y ¡x)
3
+
x ¡ y, and so on, satisfie s the given partial differential equation u
x
= u
y
= 0. This type of solution
is known as a general solution.
The concept of a general solution here is akin to the concept of the general solution for a first-
order ordinar y differential equation that typically contains a constant parameter rather than an
arbitrary function. In subsequent discussions, we will explore how to determine the specific form of
the arbitrary function h with the help of auxiliary conditions for the problem.
Exercise 2.6. Find the general solution of the equation
u
x
+ xu
y
= u:
2.2.3 Characteristic method for semi-linear PDEs
Let's consider the following equation:
v
1
(x; y)u
x
+v
2
(x; y)u
y
=v
3
(x; y; u); (2.4)
where v
1
; v
2
, and v
3
are smooth functions. The objective is to transform this partial differe ntial
equation into a set of first-order ordinary differential equations along characteristic curves.
Recall the differential of a two-variable function u = u(x; y) as du = u
x
dx + u
y
dy. Comparing
the expression of du with equation (2.4) implies the following system
du
v
3
(x; y; u)
=
dx
v
1
(x; y)
=
dy
v
2
(x; y)
: (2.5)
By relating x to y through the characteristic equation:
dy
dx
= v(x; y); (2.6)
where v =
v
2
v
1
, we obtain a family of curves γ
c
: c = g(x; y), where c is an arbitrary constant. The
given PDE reduces to t he following equation along each γ
c
:
du
dx
=
v
3
v
2
: (2.7)
Suppose this equation is solved for u = U (x; c; C), where C depends on γ
c
and thus can be expressed
as C = h(c) for an arbitrary smooth function h. Hence, the general solution can be expressed as:
u = U (x; g(x; y); h(g(x; y))):
6 Li near First-Order PDEs
Example 2.5. Let's solve the following partial differential equation:
xu
x
+ yu
y
= xy:
The characteristic equation is given by:
dy
dx
=
y
x
;
which we can solve t o obtain y = cx. The ordinary differential equat ion for u becomes:
du
dx
= y:
Substituting y = cx into this equation, we have :
du
dx
= cx:
This ordinary differential equation can be solved to find:
u =
1
2
cx
2
+ C:
Here, C can be expressed as an arbitrary smooth function in terms of c. Therefore, the gene ral
solution can be written as:
u(x; y) =
xy
2
+ h
y
x
:
Theorem 2.6. The general solution of equation ( 2.4) is
u = U (x; g(x; y); h(g(x; y))); (2.8)
where h is an arbitrary smooth function, c = g(x; y) is the equation of characteristic curves solution
of the equation ( 2.6), and U is the solution of the equation ( 2.7).
Proof. We have
(
u
x
= U
x
+ U
y
g
x
+ u
z
h
0
(g) g
x
u
y
= U
y
g
y
+ u
z
h
0
(g) g
y
:
Multiplying the first equation by v
1
and the second one by v
2
, we obtain
v
1
u
x
+ v
2
u
y
= v
1
U
x
+ U
y
(v
1
g
x
+ v
2
g
y
) + u
h
h
0
(g) (v
1
g
x
+ v
2
g
y
):
By the equation c = g(x; y), we have
0 = g
x
dx + g
y
dy;
and by the equality
dx
v
1
=
dy
v
2
;
we obtain:
v
1
g
x
+ v
2
g
y
= 0:
Hence, we obtain the equality:
v
1
u
x
+ v
2
u
y
= v
1
U
x
:
On the other hand, from the equat ion u = U (x; c; C), we have du = U
x
dx. Utilizing the equation
du
dx
=
v
3
v
1
;
2.2 Characteristic method and ODEs along curves 7
yields v
1
U
x
= v
3
, that proves the equation v
1
u
x
+ v
2
u
y
= v
3
.
Definition 2.7. Given a fixed value of c in the real numbers, t he curve γ
c
, which is the solut ion
of equation ( 2.6), is referred to as a characteristic curve of the differential equation ( 2.4). Since
equation ( 2.4) reduces to an ordinary differential equation when evaluated along γ
c
for any xed
c, we obtain an infinite system of ordinary differential equations for the family of characteristic
curves fγ
c
; c 2Rg. The system ( 2.5) is known as the characteristic system of the associated partial
differential equation.
Example 2.8. The existence of a general solution, even for linear first-order PDEs, is not always
a trivial question. Consider the equation:
xu
x
+ yu
y
= αu;
where α is a constant. We will examine three cases: α=0; ¡1, and 1.
For α=0, the characteristic system is:
dx
x
=
dy
y
=
du
0
;
and the equation for the characteristic curves in the xy-plane is:
ydx+xdy=0:
The general solution of this simple ODE is y =cx, which is shown in the following figure:
30°
60°
90°
120°
150°
180°
210°
240°
270°
300°
330°
Along the characteristic line γ
c
, we have
du
dx
=0, and thus u is constant along γ
c
. Furthermore,
all characteristic lines intersect at the origin, so γ
c
carries the information of u at the origin. This
implies that u(x; y)=u(0; 0)=C, a constant for all ( x; y). There is no other solution of the equation
in this case.
Now, consider α =¡1. We will show that the only possible solution is u0. In this case, the
solution u along γ
c
satisfies the ODE:
du
dx
= ¡
1
x
u;
and therefore, u=
C
x
along γ
c
: y =cx with respect to x. On the other hand, the PDE implies u(0;
0) =0, and thus C =0, implying that u(x; y) is identically zero in this c ase.
For α=1, the solution u satisfies the ODE:
du
dx
=
1
x
u;
with the solution u=Cx, for a constant C. The general solution in this case is:
u(x; y) = f
y
x
x:
8 Li near First-Order PDEs
The form of the solution imposes a restriction on the form of the function f if u is assumed to be
smooth inside the unit disk. For example, for f (z)=z
2
, the solution is not even continuous at the
orig in.
Exercise 2.7. Consider t he following equation
2 x
p
u
x
+ u
y
= 0:
a) Draw the characteristic curves of the equation.
b) Show that u(4; 3) = u(9; 4).
c) Write down the general solution of the equation. Note that th e solution is not differentiable with re spect
to x at x = 0.
Exercise 2.8. Consider t he following equation
u
x
+ y
p
u
y
= αu; y > 0
where α is a con stant.
a) Fi rst assume α =0. Draw the character istic curve in the xy-plane and use the relation
du
dx
= 0 to nd u(1; 1).
if u =
1
1 + x
2
on the x-axis.
b) Find the general soluti on of the equati on fo r α = ¡1. What is u(1; 1) if u =
1
1 + x
2
on the x-axis?
Exercise 2.9. Consider t he equ ation
u
x
+ u
y
= 0:
Show th at i f u is the solution of the equation, then it is impossible that u = x o n the unit circle x
2
+ y
2
= 1.
Exercise 2.10. Consider the following function
yu
x
+ xu
y
= 0:
a) What are the characteristic curves of the equation in the xy-plane?
b) Find the general soluti on of the equati on and draw some integral surfaces of the equation.
Exercise 2.11. Find the general solution of the following equations
a) xu
x
+ yu
y
= xyu:
b) ¡yu
x
+ xu
y
= 2xyu:
c) u
x
+ u
y
= x(y ¡x) u
d) xu
x
+ yu
y
= x(y + u)
Exercise 2.12. Consider the following equation
u
x
+ 3x
2
u
y
= y
1
3
a(x) + y
2
3
b(x);
where a; b are smooth functions. Show that the gener al solutio n of the equatio n is
u(x; y) = f(y ¡x
3
) A(x) + g(y ¡x
3
) B(x);
where A(x); B(x) are respectively anti-derivatives of xa(x) and x
2
b(x).
Exercise 2.13. Find the general solution of the following equation
u
x
+ u
y
= u + xu
2
:
Exercise 2.14. Consider the following equation
xu
x
+ yu
y
= f (u):
If f(0) =/ 0 , show that the equati on can not have a smooth solution inside a unit disk.
Exercise 2.15. Consider the following equation
¡yu
x
+ xu
y
= u:
If u(x; y) is a smooth solution in the closure of t he unit ball in the xy-plane , show that u = 0.
2.2 Characteristic method and ODEs along curves 9
Hint: Since u is smooth in the closure of the unit ball B, it has an absolute max and min in cl(B). If the
max and min occurs inside B, then u = 0. Let max and min occur on the boundary. Conclude the relation
¡yu
x
+ xu
y
= 0 on the boundary. Use this relation and conclude u = 0.
Exercise 2.16. Consider the following equation
yu
x
+ u
y
= u:
Take y as the independent variable for the characteristic curves x = x(y), and find the general solution of the
eq uation.
2.2.4 Semi-linear equation in general dimension
Now, we generalize the method for semi-linear PDEs in general dimension. Consider the following
equation:
X
j=1
n
v
j
(x) @
j
u = r(x; u); x 2R
n
; (2.9)
where v
j
(x) are smooth functions of x = (x
1
; : : : ; x
n
). To solve this equation, we first consider the
characteristic system:
dx
1
v
1
(x)
= ··· =
dx
n
v
n
(x)
=
du
r(x)
:
The characteristic equation in the space (x
1
; : : : ; x
n
) is
dx
1
v
1
(x)
= ··· =
dx
n
v
n
(x)
:
By taking x
1
as the independent variable, we can write the system as:
8
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
:
dx
2
dx
1
=
v
2
(x)
v
1
(x)
·
·
·
dx
n
dx
1
=
v
n
(x)
v
1
(x)
: (2.10)
We can solve system (2.10) for a given parameter c=(c
2
; : : : ; c
n
) a s the implicit system:
c
2
= g
2
(x); : : : ; c
n
=g
n
(x):
We assume that the ab ove implicit solutions are solvable for x in terms of x
1
and c, so we can write:
x
2
=x
2
(x
1
; c); : : : ; x
n
=x
n
(x
1
; c):
The intersection of surfaces
c
2
= g
2
(x); : : : ; c
n
=g
n
(x);
reduces to a smooth curve γ
c
, c=(c
2
; : : : ; c
n
), which is the same as the parametric curves:
γ
c
: fx
2
=x
2
(x
1
; c); : : : ; x
n
=x
n
(x
1
; c)g:
Next, we consider the derivative of u along the curve γ
c
, which is given by:
du
dx
1
=
r(x; u)
v
1
(x)
: (2.11)
By solving this equation, we obtain a function u = U (x
1
; c
2
; : : : ; c
n
; C). Therefore, taking C =h(c),
for an arbitrary smooth function f , the general solution to equation (2.9) can be expressed as:
u = U (x
1
; g
2
(x); : : : ; g
n
(x); h(g
2
(x); : : : ; g
n
(x))):
10 Li near First-Order PDEs
Example 2.9. Let's solve the partial differential equation:
u
x
+ u
y
+ u
z
= 0:
The characteristic equations in the space (x; y; z) is given by:
dy
dx
= 1;
dz
dx
= 1;
which is solved for c
2
= y ¡x, c
3
= z ¡x. The characteristic curves parameterized by x is
γ
c
= (x; x + c
2
; x + c
3
)
for c = (c
2
; c
3
). The given PDE along γ
c
b ecomes:
du
dx
= 0, which implies u=C. Replacing C by
h(c
1
; c
2
) for an arbitrary smooth function h, we obtain the general solution
u(x; y; z)=h(y ¡x; z ¡x):
Exercise 2.17. Consider the following equation
u
x
¡zu
y
+ yu
z
= αu;
where α is a con stant.
a) Show that the characteristic curves of the equation have the following equat ions
γ
fc
1
;c
2
g
(x) = (x; c
1
cos(x) ¡c
2
sin(x); c
1
sin(x) + c
2
cos(x)):
What is the shape of the characteristic curves?
b) The characteristic curves passes through the plane x = 0. Show that for α = 0, we have
u(0; 1; 1) = u
π
2
; ¡1; 1
:
c) Show that the solution of the given PDE along γ
fc
2
;c
2
g
(x) is of the form
u(x; c
1
; c
2
) = h(c
1
; c
2
) e
αx
;
for arbitrary function h. Find t he general solution.
Exercise 2.18. Consider the following equation
u
x
+ 2 y
p
u
y
+ 2 z
p
u
z
= 0;
where y 0 and z 0. Find the general solution of the equation. Veri fy the equ ality u(2; 4; 4) = u(1; 1; 1).
Exercise 2.19. The singular equation
xu
x
+ yu
y
+ zu
z
= 0:
If u(x; y ; z) is a smooth solution inside the unit ball in R
3
, show that u(x; y; z) = C a constant for all x; y ; z.
2.2.5 Characteristic method for quasi-linear equat ions
The method of characteristics for quasi-linear partial differential equations may lead to the emer-
gence of new phenomena, such as shocks, which we will discuss later. Let's consider the following
equation
v
1
(x; y; u) u
x
+ v
2
(x; y; u) u
y
= v
3
(x; y; u): (2.12)
The coefficients of the quasi-linear equations depend on u. The characteristic system of the PDE is:
dx
v
1
=
dy
v
2
=
du
v
3
: (2.13)
2.2 Characteristic method and ODEs along curves 11
Let c
1
= φ(x; y; u) and c
2
= (x; y; u) be implicit solutions of the characteristic system. We have
the following theorem:
Theorem 2.10. Let v
1
;v
2
, and v
3
be smooth functions in equation ( 2.12), and let c
1
=φ(x; y;u) and
c
2
= (x; y; u) be implicit solutions of its associated characteristic system. Then, the general solution
of the equation in implicit form is given by f(φ; )=0 for any smooth function f satisfying f
u
=/ 0.
Example 2.11. Let's consider the following equation called the Burger's equation:
u
x
+ uu
y
= 0:
The characteristic system is given by:
dx
1
=
dy
u
=
du
0
;
By taking x as the independent variable and setting y = y(x), we obtain:
dy
dx
= u;
du
dx
= 0:
This system is solved for c
1
= u and c
2
= y ¡xu. The general implicit solution of the equation is
derived by f(c
1
; c
2
) = 0, for arbitrary smooth function f, or equivalently f(u; y ¡xu) = 0. Alterna-
tive ly, we can write the solution as u = g(y ¡xu) for arbitrary smooth function g. Note that u in
both cases is in the implicit form.
Example 2.12. Consider the equation
u
x
+ yuu
y
+ zuu
z
= 0:
To solve this equation, we need to determine the characteristic system, which is given by:
dy
dx
= yu;
dz
dx
= zu;
du
dx
= 0:
Solving this system of differential equations, we obtain e
¡x
y = c
1
u, e
¡x
z =c
2
u, and u=c, where c;
c
1
and c
2
are arbitrary constants. Substituting these values into the equation, we get the general
implicit solution as
f(u; e
¡x
yu
¡1
; e
¡x
zu
¡1
) = 0:
Alternatively, the solution can be expre ssed as u = g
ye
¡x
u
;
ze
¡x
u
.
Proof. (of the theorem) Since f (φ(x; y; u(x; y)); (x; y; u(x; y))) is identically zero, taking
derivatives with respect to x and y gives:
(
f
φ
(φ
x
+ u
x
φ
u
) + f
(
x
+ u
x
u
) = 0
f
φ
(φ
y
+ u
y
φ
y
) + f
(
y
+ u
y
u
) = 0
:
Note that f
u
=/ 0 implies that f
φ
and f
cannot both be identically zero. This in turn implies that
the following determinant is zero:
φ
x
+ u
x
φ
u
x
+ u
x
u
φ
y
+ u
y
φ
y
y
+ u
y
u
= 0:
Expanding the determinant and simplifying, we get:
(φ
u
y
¡
u
φ
y
)u
x
+ (φ
x
u
¡ φ
u
x
) u
y
= φ
y
x
¡ φ
x
y
: (2.14)
12 Li near First-Order PDEs
On the other hand, we have:
(
= v
1
φ
x
+ v
2
φ
y
+v
3
φ
u
=0
d = v
1
x
+ v
2
y
+ v
3
u
= 0
;
that gives
v
1
φ
u
y
¡
u
φ
y
=
v
2
φ
x
u
¡ φ
u
x
=
v
3
φ
y
x
¡φ
x
y
: (2.15)
Matching relations (2.14) and (2.15) gives v
1
u
x
+ v
2
u
y
= v
3
, which completes the proof.
Exercise 2.20. Find the general solution of the following equations
a) uu
x
+ u
y
= u
2
b) sinuu
x
+ u
y
= 1 (hint: you can take y a s the independent variable just for the simplicity)
c) xu
x
+ (1 + u) u
y
= u
d) u
x
+ uu
y
¡uu
z
= 1
Exercise 2.21. Consider the following equation
u
x
+ uu
y
= 0:
a) Fi nd the general sol ution in the implicit fo rm. Then for the auxiliary condition u(0; y) = y, find an explici t
solution. Determine the dom ain (x; y) where th e solution exists smoothly.
b) In general, the im plicit solutions can not be transformed to an explicit form, although, th e exis tence of
such explicit forms is guaranteed by the implicit function th eorem. Let u(0; y) = e
y
. Obtain the implicit
solution of the equation and use the theorem to show the existence of the expl icit solution.
Exercise 2.22. Consider the following equation
(
u
x
+ uu
y
+ uu
z
= 0
u(0; y; z) = f (y; z)
:
a) Fi nd the solution of the equat ion in the implicit form.
b) Find the explicit solution if f(y; z) = y + z.
Exercise 2.23. Consider the following quasi-linear equation
@
1
u +
X
j=2
n
v
j
(x; u) @
j
u = r(x; u); x 2R
n
The cha racteristic systems is
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
dx
j
dx
1
= v
j
; j = 2; : : : ; n
du
dx
1
= r
:
Let the system sis solved fo r
c = (x; u); c
j
= φ
j
(x; u); j = 2; : : : ; n:
Show th at the function f (φ
j
(x; u); (x; u)) = 0 solves the given partial differential equa tion.
2.3 Theoretical aspects
2.3.1 The geometrical interpretation of a first-order PDE
Consider the partial differential equation:
v
1
(x; y; u) u
x
+ v
2
(x; y; u) u
y
= v
3
(x; y; u): (2.16)
2.3 Theoretical aspects 13
Here, v
1
; v
2
, and v
3
are functions of x; y, and u. Let z = u(x; y) is a solution to this equation. This
function defines a surface on the space (x; y; z):
S: (x; y; u(x; y)):
Recall that the vector n~ : (¡u
x
; ¡u
y
; 1) is perpendicular on S at any point (x; y; z) on S. Now,
assume that the vector field V
~
: (x; y; z) 7!(v
1
; v
2
; v
3
) is given on the space. The partial differential
equation (2.16) then can be stated from the geometrical point of view as follow:
Geometrical interpretation of a first-order PDE: Given a vector field V
~
: (x; y; z)!(v
1
;
v
2
; v
3
), the equation of a tangent surface z = u(x; y) to V
~
is given by the partial differential equation
v
1
(x; y; u) u
x
+ v
2
(x; y; u) u
y
= v
3
(x; y; u):
Conversely, the solution of the above partial differential equation defines a surface z = u(x; y) that
is locally tangent to the vector field V
~
.
For example, consider the equation: u
x
+ u
y
= 0. The vector field is
V
~
: (x; y; z) !(1; 1; 0)
The figure below shows two different surfaces that are tangent to this vector field at all point of the
surfaces
In fact, it can be seen that the all surface generated by the function u = f (y ¡x) for arbitrary
smooth f has this tangent property as
V
~
·n~ = (1; 1; 0) ·(f
x
0
; ¡f
y
0
; 1) = 0 :
Problem 2.1. Let V = (v
1
(x; y; z); v
2
(x; y; z); v
3
(x; y; z)) be a given vector field in space.
a) Show that the existence of a surface φ(x; y; z) = 0 which is perpendi cular to the vector fields at each point
on the surface satisfies the following equation
v
1
dx + v
2
dy + v
3
dz = 0:
b) The exi stence of the solu tion to the a bove ODE is not a trivia l questi on. Show that if r×V = 0, then there
is a sur face φ(x; y; z) = 0 that sa tisfies the above ODE. Hint: Note that if r×V = 0, then V is potential.
c) As an examp le, consider the vector eld V = (y; x; z). Find a surface which is perpendicular to V at each
point on the surface. The gure below depicts one of such surfaces .
14 Li near First-Order PDEs
Problem 2.2. Consider the vector field V = (¡y; x; 0). This field does not sati sfy the condition r × V = 0,
however, a su rface still exists which is per pendicular to V . Show that the surfa ce expressed pa ra metrically as
Σ(x; z) = (x; αx; ¡xz) has such a property.
Problem 2.3. Show that there is no surface φ(x; y; z) = 0 with the normal vectors (x; y; xy).
2.3.2 Parametric solution surfaces
Now, consider a curve γ(t) on the solution surface (yet unknown) S. This curve is tangent to the
vector field V
~
at any point on the curve. Therefore, the equation of γ(t) is determined by the
following ordinary differential equation
8
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
:
dx
dt
= v
1
(x(t); y(t); z(t))
dy
dt
= v
2
(x(t); y(t); z(t))
dz
dt
= v
3
(x(t); y(t); z(t))
: (2.17)
This first-order system of ordinary differential equation has a unique solution if an initial condition
is set for the system. Let's assume that we know a point p
0
on S. We can set the initial condition as:
γ(0) = p
0
:
8
>
>
<
>
>
:
x(0) = x
0
y(0) = y
0
z(0) = z
0
:
With this initial condition, a curve γ
p
0
(t) is obtained on S. In this way, to determine S, we need
to a family of curves fγ
p
(t)g where p lies on a curve on S as shown below:
γ
p
(t) = Σ(t; s)
z = u(x; y)
x
p(s)
y
For example, assume the curve p(s) = (s; 0; s sin(s)) lies on S o f the solution of the equation
u
x
+ u
y
= 0. The solution surface is spanned by the curves det ermined by the following system of
ODEs
8
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
:
dx
dt
= 1
dy
dt
= 1
dz
dt
= 0
;
accompanied with the conditions x(0) = s; y(0) = 0; z( 0) = s sin(s). This system is solved as
x(t; s) = t + s; y(t; s) = t; z(t; s) = s sin(s):
2.3 Theoretical aspects 15
In this way, we obtain a parametric surface
Σ(t; s) = (t + s; t; s sin(s)):
It is simply seen that this parametric surface is algebraically represented as
u(x; y) = (y ¡x) sin(y ¡x):
The figu re below depicts the “data line” p(s) in black and the space characteristic curves γ
p
(t) in red.
Remark. The advantage of using parametric form for representing solutions is that it can represent
very complicated surfaces, whereas explicit functions z = u(x; y) represent only restricted classes of
surfaces. The example below furt her clarifies this point .
Example 2.13. Consider the following problem
(
0.2xu
x
¡uu
y
= y
uj
fx=y
2
g
=y
:
Here u is give n along the curve x = y
2
in the xy-plane. This data can be parameterized in terms of
s as: p(s)=(s
2
; s; s). The system of characteristic equations is:
8
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
:
dx
dt
= 0.2x
dy
dt
= ¡u
dz
dt
= y
;
with the initial con ditions x
0
= s
2
; y
0
= s and z
0
= s. The parametric representation of the integral
surface is obtained as:
Σ(t; s) = (s
2
e
0.2t
; s(cost ¡sint); s(sint + cost)):
As shown in figure below, the solution of the equation represents a complicated surface which can
not be expressed by an explicit function z = u(x; y).
16 Li near First-Order PDEs
Even though this surface is not associated with an explicit function, it is possible to define
an explicit function that locally coincides with this surface. Thus, classical solutions of partial
differential equati ons c an only be defined locally in a neighborhood of the data curve.
Exercise 2.24. Use the parametrization and find the solu tion of the following Ca uchy problems. Use Matlab
and draw the integral surface
a) uu
x
+ yu
y
= x, u = sin(x) on y = 1.
b) (y ¡u) u
x
+ (u ¡x) u
y
= u, u = x
2
on y = 0.
The following code generates the integral surfaces of t he given equation in part b).
s = lins pace(0, 2, 100); % d efine the range of s
t = lins pace(0, 10, 100); % define the range of t
[S, T] = meshgrid(s, t); % create a grid of s and t values
X = S .* cos(T); % compute the x-coordinates
Y = S .* sin(T); % compute the y-coordinates
Z = S .^ 2 .* exp(0.2 * T);% compute the z-coordinates
surf(X, Y, Z) % plot the surface
shading interp; camlight
xlabel('x');
ylabel('y');
zlabel('z');
title('Surface plot of x = s*cos(t), y = s*sin(t), z = s^2*exp(0.2*t)');
2.3.3 Cauchy problem
In this section, we demonstrate how to obtain the particular solution o f a given first-order PDE
from the general solution, using an auxiliary condition or additional information about the solution.
This process is similar to deriving particular solutions from the general solutions of ODEs, using
initial conditions.
Definition 2.14. A problem of the form
(
v
1
(x; y; u) u
x
+ v
2
(x; y; u) u
y
= v
3
(x; y; u)
uj
C
=g
(2.18)
where C is a curve in an open set R
2
in the xy-plane, is called a Cauchy problem.
Let's consider the following problem:
8
<
:
u
x
+ u
y
= 0
uj
fy=0g
=
1
1 + 2x
2
:
The general solution to the equation is u = h(y ¡x) for an arbitrary smooth function h. Utilizing
the initial condition u(x; 0) =
1
1 + 2x
2
is
1
1 + 2x
2
= h( ¡x);
and thus h(y ¡x) =
1
1 + 2(y ¡x)
2
, and thus t he particular solution to the given Cauchy problem is
given by:
u(x; y) =
1
1 + 2(y ¡x)
2
:
2.3 Theoretical aspects 17
Exercise 2.25. Find the partic ular solution of the followi ng Cauchy problem
(
u
x
+ (x + y)u
y
= u
uj
x=0
=sin(y)
:
Exercise 2.26. Find the partic ular solution of the followi ng Cauchy problem
(
u
x
+ u
y
= 2x(y ¡x)u
uj
x=0
=e
y
:
Exercise 2.27. The Cauchy problem can be gener alized to higher dimensi ons. Fi nd the solution of the following
eq uation for u = u(x; y; z):
(
xu
x
+ yu
y
+ z u
z
= 0
uj
fz=1g
=xy
Examples below highlight some of the issues that can arise when attempting to extract a par-
ticular solution from a general solution.
Example 2.15. Consider the Cauchy problem:
(
u
x
¡2x u
y
= 0
uj
fy=x
2
g
=x
:
The general solution of the PDE is u = f (y + x
2
) for an arbitrary smooth function f. We can use the
auxiliary condition u = x on the line C: y = x
2
to determine f , which gives us x = f(2x
2
). However,
this equation cannot be solved as f ( 2 ) can have two possible values, ±1, leading to a contradiction.
-2 -1 1 2
-4
-2
2
4
This problem occurs because the characteristic curves of the PDE, y =¡x
2
+c, intersect the data
curve C: y = x
2
at more than one point, resu lting in multiple possible values of u. To resolve this
issue, we can consider only one branch of the data curve, such as y = x
2
; x 0, to obtain a unique
solution u(x; y) =
y + x
2
2
q
for y ¡x
2
.
-2 -1 1 2
-4
-2
2
4
However, this solution is still not unique in the region y < ¡x
2
, where the characteristic curves
do not intersect the data curve. The value of u along these curves can be chosen arbitra rily.
18 Li near First-Order PDEs
-2 -1 1 2
-4
-2
2
4
If we change the data curve to y = ¡x
2
, which is also a characteristic curve, there is no unique
solution in the region y > ¡x
2
and y < ¡x
2
. In this case, x = f (0), which i s not solvable from the
general solution u = f (y + x
2
) and the data curve y = ¡x
2
.
-2 -1 1 2
-4
-2
2
4
Example 2.16. Let us consi der the following equation:
(
¡yu
x
+ xu
y
= u
uj
C
=x
;
where C is the x-axis f or x 0. To apply the method of characteristics, we first need to find the
equation of the characteristic curves. Using the characteristic equation
dx
¡y
=
dy
x
;
we get x
2
+ y
2
= c; where c is a positive constant. Therefore, the characteristic curves are circles
centered at the origin. In one of previous exercise, we asked the read to show that this equation does
not have any smooth solution inside a disk. We can see this fact by trying to solve this equation
explicitly. Taking x as the independent variable, we can rewrite the PDE as
du
dx
=
¡1
y
u:
To solve this equation, we need to express y as a function of x. Howe ver, this cannot be done
through the obtained implicit function x
2
+ y
2
= c. One way to overcome this difficulty is to use
the parametric representation of the characterist ic curves. Let t be a parameter and consider the
characteristic system:
8
<
:
dx
dt
= ¡y
dy
dt
= x
:
The solution to this system is
γ
p
(t) = (s cos(t); s sin(t));
2.3 Theoretical aspects 19
where s is a non-negative parameter. Note that we used the data curve C to write the initial point
of the cha racteristic curve as γ
s
(0) = (s; 0). Now, we can express u in terms of t as f ollows:
du
dt
= u;
which is a separable ODE with solution
u(γ
s
(t)) = u(γ
s
(0)) e
t
= u(s) e
t
= se
t
:
To determin e the domain o f t, note that γ
s
(0) = γ
s
(2π), which implies
u(γ
s
(2π)) = u(γ
s
(0)) = s:
However, we also have
u(γ
s
(2π)) = se
2π
;
so we conclude that the domain of t can not be [0; 2π]. Note that for x = s cost; y = s sint and u = se
t
,
we obtain the integral surface
Σ(t; s) = (s cost; s sint; se
t
):
This solution can be put in the algebraic form as
u(x; y) = x
2
+ y
2
p
e
atan
¡
y
x
:
Exercise 2.28. Find the partic ular solution of the followi ng Cauchy problem
(
u
x
+ (x + y)u
y
= u
uj
x=0
=sin(y)
:
Exercise 2.29. Find the partic ular solution of the followi ng Cauchy problem
(
u
x
+ u
y
= 2x(y ¡x)u
uj
x=0
=e
y
:
Exercise 2.30. The Cauchy problem can be gener alized to higher dimensi ons. Fi nd the solution of the following
eq uation for u = u(x; y; z):
(
xu
x
+ yu
y
+ z u
z
= 0
uj
fz=1g
=xy
Exercise 2.31. Consider the following Cauchy problem
(
¡yu
x
+ !
0
2
u
y
= αu
uj
C
=x
;
where C is the line y = x for x 0, and !
0
is a constant. Use parametric representation of characteristic curves in
ter ms of the parameter t and nd the solution for α = 0. If α =/ 0 , nd the domain of t and indica te the d omain
of the soluti on in the xy-plane.
Exercise 2.32. Consider the following Cauchy problem
(
(¡y ¡αx) u
x
+ (x ¡αy) u
y
= u
uj
C
= f (x)
;
where C is the x-axis for x 0, and α > 0. Show that the problem is solvable in classical sen se only if f(x) = 0.
2.3.4 Well-posedness and existence of i ntegral surfaces
As we observed in previous examples, if the data curve of a Cauchy problem is not a characteristic
curve, then there is a solution that can be extended locally. The following figure shows this situation
schematically.
20 Li near First-Order PDEs
γ
s
0
(t)
C(s)
m
1
m
2
x
C(s
0
)
y
Here, the data cur ve C is parametrized by s as C = C(s), and the planar characteristic curves
γ are parameterized by t. Note tha t γ
s
(t) is a characteristic curve passing throu gh C(s) at t = 0.
Let s
0
b e a point on C(s) in its domain. If γ
s
0
0
(0) and C
0
(s
0
) are non-parallel, then there is a t
0
> 0
such that γ
s
0
(t) exists for 0 t < t
0
.
Theorem 2.17. Consider the Cauchy problem
(
v
1
(x; y) u
x
+ v
2
(x; y) u
y
= v
3
(x; y; u)
uj
C
=f
where v; v, and v are smooth functions, and C is a smooth curve in the xy-plane. Assume there
exists (x; y)2C such that
C
0
(x
0
; y
0
) jj (v
1
(x
0
; y
0
); v
2
(x
0
; y
0
)): (2.19)
Then there exists an open neighborhood of (x; y) and a smooth function u = u(x; y) on that
solves the given Cauchy problem.
The p roof of the theorem is based on a standard theorem on the existence and uniqueness of
the solution to ordinary differential equations. Note that if condition (2.19) holds, then due to the
continuity of v
1
; v
2
, and C
0
at (x
0
; y
0
), the condition holds for an open neighborhood of (x
0
; y
0
).
Then, the existence of a domain for the solution u(x; y) is reduced to the existence and uniqueness
of the ordinary differential equation
du
dt
= v
3
(x(t); y(t); u):
However, it is important to note that the theorem only provides a sufficient condition for the
existence of a solution, and there may be cases where the condition is not satisfied but a solution
still exists. For instance, consider the problem:
(
u
x
+ y
p
u
y
= 0
u(x; 0) = f(x)
This problem has the solution u(x; y) = f(x ¡2 y
p
) which is defined for y 0, even though it i s not
generally differentiable on the x-axis.
The existence of a parametric surfac e for a quasi-linear first-order PDE is established in a similar
manner. We have the following theorem.
Theorem 2.18. Consider the Cauchy problem
(
v
1
(x; y; u) u
x
+ v
2
(x; y; u) u
y
= v
3
(x; y; u)
uj
C
=f
;
2.3 Theoretical aspects 21
where v
1
; v
2
; v
3
are smooth functions and C is a smooth and non-characteristic curve in xy-plane.
Let ¡(s) be the parametrized space data curve (C(s); f(s)). Fix s
0
in the domain of ¡(s) and let
(x
0
; y
0
; u
0
) = ¡(s
0
). If
¡
0
(s
0
) ×V (x
0
; y
0
; u
0
) =/ 0;
where V = (v
1
; v
2
; v
3
), then there exist α; β >0 such that the given problem has an integral surface
Σ(t; s) for t 2(¡β; β) and s 2(s
0
¡α; s
0
+ α).
2.4 Time de pendent functions and fluid flow
Consider a fluid distributed along the x-axis at time t, with u(x; t) representing its density function.
Let x
0
b e a fixed point on the x-axis. The rate of change in u(x
0
; t) as time progresses is given
by the partial derivative u
t
(x
0
; t). Meanwhile, u
x
(x; t
0
) measures the rate of difference in density
between x and its neighboring points.
x
t
(x
0
; t
0
)
x
x
t
t
(x
0
; t
0
)
u
t
(x
0
; t
0
)
u
x
(x
0
; t
0
)
Now, consider a fluid with density function u(x; t) owing with velocity function V (x). This
veloc ity function can be thought of as a one-dimensional vector field: x!V (x). The properties of
this field affect the density function u. For example, suppose we have a control volume [¡1; 1 ] moving
with velocity V (x)=x. The total mass within this volume at time t = 0 is given by the integral
M =
Z
¡1
1
u(x; 0) dx:
What can be said about the mass of this packet at time t = 1? Let x
0
b e a fixed point on the x-axis.
According to the velocity field, the p osition of x
0
at time t can be deter mined by the differential
equation
dx
dt
=V . For V (x)=x, we obtain x(t)=x
0
e
t
. Therefore, the control volume [¡1; 1] expands
to the volume [¡e; e] at time t = 1. Assuming there is no sink or source of mass production, the
total mass is conserved, and thus
M =
Z
¡e
e
u(x; 1) dx:
Since the volume of the control volume has increased, the density u(x; 1) must have decreased to
maintain the same mass. The main objective of this section is to address how the density fu nction
u of a n-dimensional fluid flow changes over time.
Exercise 2.33. Consider the following density distr ibution
u(x; t) = e
¡
(x¡ t)
2
2
:
It is simp ly seen that the total mass al ong the x-axis for any t is equal to
Z
¡1
1
u(x; t) dx = 2π
p
;
22 Li near First-Order PDEs
and thus we can write that
lim
t!1
Z
¡1
1
u(x; t) dx = 2π
p
:
Find the steady state distribution when t !1 and calculate
Z
¡1
1
lim
t!1
u(x; t) dx:
This shows that the foll owing eq uality does not hold always
lim
t!1
Z
¡1
1
u(x; t) dx =
Z
¡1
1
lim
t!1
u(x; t) dx:
2.4.1 Change of density along flow lines
Consider a velocity fiel d V = (v
1
(x); : : : ; v
n
(x)) of a fluid flow, where x = (x
1
; : : : ; x
n
). Let x
0
b e a
fixed point in R
n
, and let γ
x
0
(t) be the flow line of x
0
, determined by the system of ODEs
dx
j
dt
= v
j
(x); j = 1; : : : ; n;
with initial condition x(0)=x
0
. Our goal is to find the rate of change of the density function u along
the flow line γ
x
0
(t). This rate of change is given by the time derivative:
du
dt
(γ
x
0
; t) = u
t
(γ
x
0
; t) + ru(γ
x
0
; t) ·
dγ
x
0
dt
;
where ru(t; x) represents the gradient of u. This can be expressed in coordinate form as:
du
dt
(x; t) = u
t
(x; t) +
X
j=1
n
v
j
(x) @
j
u(x; t):
The derived formula provides us with the change of u along the flow lines. Therefore, the question
arises: can we determine u(x; t) if we know the initial density u(x; 0)?
Exercise 2.34. Verify that the above formula is a straightforward applica tion of the chain rule.
Example 2.19. Consider a fluid flow in the xy-plane, where the density at t = 0 is given by
u(x; y; 0) = f(x; y) = x
2
e
¡x
2
¡y
2
, and the velocity field V is
(x; y) 7!(¡αx ¡y; x ¡αy);
with α 0 being a constant. The trajectories of the particles are determined by the following system
of differential equations:
8
<
:
dx
dt
= ¡αx ¡ y
dy
dt
= x ¡αy
:
For the first case, let α = 0, that leads to the trajectory
γ
(x
0
;y
0
)
(t) = (x
0
cos(t) ¡ y
0
sin(t); x
0
sin(t) + y
0
cos(t));
where (x
0
; y
0
) is the initial position at t = 0. In matrix form, the flow line passing through (x
0
; y
0
)
is given by:
x(t)
y(t)
=
cos(t) ¡sin(t)
sin(t) cos(t)
x
0
y
0
:
2.4 Time dependent functions and fluid flow 23
Note that the coefficient matrix is the rotation matrix R
t
. Consider a control volume D centered
at (0.5; 0.5). The following graph illustrates u(x(t); y( t); t) for t = 0 :
π
4
: 2π. Observe how the control
volume centered at (0.5; 0.5) rotates around the origin while also rotating around itself. The velocity
field is responsible for this rotation, as it is a rotational field with a non-zero curl, denoted as
r×V = 2.
The following figure shows the density at different time slices, specifically for times t = 0 ;
π
2
, and
π. As shown in the figure, the density rotates in a circular pattern around t he origin as it is clear
from the differential equations of the trajectory
dx
dt
= ¡y;
dy
dt
= x. Furthermore, we observe that the
density along the trajectory of the fluid particles remains constant, which means the density remains
constant a long the trajectory. This type of fluid flow, where the density remains constant along the
trajectory of the fluid particles, is called an incompressible flow.
Now, let α = 0.1. The trajectory of the control volume is an inward spiral towards the origin,
given by
x(t)
y(t)
= e
¡0.1t
cos(t) ¡sin(t)
sin(t) cos(t)
x
0
y
0
:
The following graph illustrates this scenario, showing how the control volume centered at (0.5; 0.5)
forms an inward spiral while also rotating around itself. Observe how the control volume gradu ally
becomes smaller and by the conservation law, the density in D increases to balance the decrea se in
the volume.
24 Li near First-Order PDEs
The above example illustrate how the velocity filed V ( x) affects the density of particles along
the flow lines. The following theorem make a relationship between the divergence r·V (x) and the
change of density along the flow line.
Theorem 2.20. Consider a control volume D in a fluid flow that moves according to the velocity
field V. If the divergence of V, denoted by r·V (γ
D
(t)), is positive, then the volume of D increases
along the trajectory γ
D
(t). On the other hand, if r·V (γ
D
(t)) is negative, the volume of D decreases
along the trajectory, and if r·V (γ
D
(t)) is zero, the volume of D remains constant.
Corollary 2.21. If r·V = 0, the density of a fluid-flow remains constant along its trajectory in
the absence of any material sink or source.
In the first scenario of the example, the velocity field V = (¡y; x) had a divergence of r·V = 0,
which implies that the volume of the control volume remained constant along γ
D
(t). As a result,
the density u remained unchanged along the flow lines. In the second scenario, the velocity field
was V = (¡y ¡αx; x ¡αy), with a divergence of r·V = ¡2α < 0 for α > 0. This indicates that the
volume of the control volume decreased with time, leading to an increase in the density u to balance
the volume decrease.
2.4.2 Continuity equation
Given a region in a fluid flow, the total mass of the fluid in at time t, denoted by M
(t), changes
due to two factors. The first factor is the net flow rate of fluid passing through the boundary bnd(Ω)
per unit time. This can be expressed mathematically as the surface integral
Φ(t) := ¡
ZZ
bnd(Ω)
u(x; t)V (x) ·νdV :
where V is the velocity vector of the fluid and ν is the unit outward normal vector to the boundary.
The second factor is the presence of a source or sink of material in the fluid, which can be represented
mathemat ically as the volume integral
Q(t) =
ZZZ
f(x; t) dV ;
2.4 Time dependent functions and fluid flow 25
where f (x; t) is the rate of material added or removed per unit volume at point x at time t.
Therefore, we can write the rate of change of M
(t) as:
dM
(t)
dt
= ¡
ZZ
bnd(Ω)
u(x; t)V (x) ·νdV +
ZZZ
f(x; t) dV :
This equation is known as the mass conservation equation, or the continuity equation, and it plays
a fundamental role in fluid dynamics.
On the other hand, the left hand side of the above equation is
dM
(t)
dt
=
d
dt
ZZZ
ρ(x; t) dV =
ZZZ
ρ
t
(x; t) dV ;
and by equating the two expressions for
dM
dt
, we arrive at the continuity equation for the fluid flow
in :
ZZZ
ρ
t
(x; t) dV = ¡
ZZ
bnd(Ω)
u(x; t)V (x) ·νdS+
ZZZ
f(x; t) dV :
The material flux J(x; t) = u(x; t)V (x) measures the rate at which fluid passes through a unit area
per unit time. When considering the flux at a particular point x on the boundary bnd(Ω) of the
bounded, open set in R
3
, the unit normal vector ν(x) is needed to determine the flux through the
surface. This is because the flux J(x; t) can be decomposed into two components: 1) a component in
the direction of the unit normal vector ν(x), and 2) a tangential component that does not leave the
closure of , denoted by
¯
. Only the component of J(x; t) that is perpendicul ar to the boundary,
i.e., J(x; t) ·ν(x), contributes to the flow leaving through bnd(Ω). The figure below illustrates
this concept.
The differential form of the continuity equation can be derived by applying Gau ss's theorem to
the divergence of the material flux density. Specifically, the theorem yields:
ZZZ
u
t
(x; t) dV = ¡
ZZZ
r·[u(x; t)V (x)]dV +
ZZZ
f(x; t) dV ;
which simplifies to:
ZZZ
fu
t
(x; t) + r·[u(x; t) V (x)] ¡ f (x; t)gdV = 0:
Since the integral holds for every arbitrary subdomai n of , we conclude that the integrand must
be identically zero, yielding the continuity equation:
u
t
(x; t) + r·[u(x; t) V (x)] = f ( x; t): (2.20)
This equation expresse s the conse rvation of mass, where f(x; t) represents the source or sink of
material i n the fluid.
26 Li near First-Order PDEs
Exercise 2.35. Show the relation
div [u(x; t) V (x)] = ru ·V (x) + u(x; t) div [V (x)];
and conclude that t he conservation equation can be written as follows
u
t
+ V ·ru = ¡u div [V ]:
Assuming f is identically zero in equation (2.20), we can rewrite the equation using the result
from the previous exercise as follows:
u
t
+ V ·ru = ¡ur·V : (2.21)
Let γ
z
(t) be the trajectory of a particle at z which is determined by the equatio n
dz
dt
= V (z). In this
case, the left hand side of the above equation reads
du
dt
(t; γ
x
(t)), and then we have
du
dt
(γ
z
(t); t) = ¡ur·V (γ
z
(t)):
If r·V (γ
z
(t)) = 0, then
du
dt
(γ
z
(t); t) and this implies that the density of x remains the same a t its
trajectory, an therefore, equation ( 2.21) leads to the solution
u(γ
z
(t); t) = u(γ
z
(0); 0);
and by the fact γ
z
(0) = z, we obtain
u(γ
z
(t); t) = u(z; 0):
If we denote γ
z
(t) by x and then z = γ
x
(¡t), we derive the solution
u(x; t) = u(γ
x
(¡t); t):
Definition 2.22. A flow moving along the velocity filed V (x) is called incompressible if div (V ).
Exercise 2.36. We aim to show th at the total mass within a moving c ontrol volume D, defined by the trajectory
γ
D
(t), remains constant when there are no external sources or sinks. To begin, we ex press the total mass M
D(t)
as the integral of the flui d density u(x; t) over the volume D(t) at time t:
M
D(t)
=
ZZZ
D (t)
u(x; t) dx:
Taking the time derivative of the above equation, we get:
dM
D (t)
dt
=
d
dt
ZZZ
D(t)
u(x; t) dx:
Using the Leibniz integral rule, we can write:
d
dt
ZZZ
D (t)
u(x; t) dx =
ZZZ
D(t)
u
t
(x; t) dx +
ZZ
bnd(D(t))
uV (x) ·ν dS:
Use the divergence theorem (Gauss theorem) and show
d
dt
ZZZ
D (t)
u(x; t) dx = 0;
and conclude that M
D (t)
remains constant along γ
D
(t).
2.5 Linear an d semi-linear transport equations
The general form of a semi-linear transport equation in R
n
is given by
u
t
+V (x) ·ru= f (x; t);
2.5 Li near and semi-linear transp ort equatio ns 27
where V (x) is the velocity field. However, in some cases, the velocity field may also depend on the
density u in addition to x, leading to the quasi-linear equation:
u
t
+V ( x; u) ·ru= f(x; t):
In general, the independent variable for transport equations is time t. For semi-linear equations,
the characteristic system is given by
dx
dt
=V (x);
while for quasi-linear equations, the characteristic system becomes:
8
<
:
dx
dt
= V (x; u)
du
dt
= f(x; t)
:
Example 2.23. The simplest form of the transport equation is:
u
t
+ c u
x
= 0
u(x; 0) = u
0
(x)
:
This equation is also known as the unilateral wave equation due to t he directionality of the wave
propagation. Since V (x)=c is constant, the flow is incompressible, and the trajectory x(t)=ct+x
0
yields the ordinary differential equation
du
dt
= 0, which has the solution:
u(x(t); t)=u(x
0
; 0)=u
0
(x
0
) = u
0
(x ¡ct):
This solution corresponds to a traveling wave propagating to the right (if c>0) or to the left (if c<0 ).
The solution to the transport equation in h igher dimensions is similar. Consider the equation:
u
t
+ C ·ru = 0
u(x; 0) = u
0
(x)
;
where C = (c
1
; :::; c
n
) and x = (x
1
; : : : ; x
n
). The flow lines are:
x
j
(t) = c
i
t + x
0
j
;
where j =1; :::; n and x
0
j
= x
j
(0). Thus, we have u(t; x(t))= f(x
0
), and for x
0
= x ¡Ct, the solution is:
u(x; t) = u
0
(x
1
¡c
1
t; : : : ; x
n
¡c
n
t):
Example 2.24. Consider the following equation
8
<
:
u
t
+ cxu
x
= αu
u(x; 0) =
1
1 + x
2
;
Here, c is a constant. When α = ¡c, the equation is conservative, and can be written as:
u
t
+ r·[cxu] = 0:
Let's first solve the equation for α = ¡c, where c > 0. We can derive the flow lines by solving:
dx
dt
= cx;
28 Li near First-Order PDEs
which gives us γ
x
0
: x(t) = x
0
e
ct
. The PDE along γ
x
0
can then be expressed as:
du
dt
= ¡cu;
which can be solved to give:
u(x(t); t) = u(x
0
; 0) e
¡ct
=
e
¡ct
1 + x
0
2
:
Replacing x
0
by xe
¡ct
, we obtain
u(x; t) =
e
¡ct
1 + x
2
e
¡2ct
:
As expected, the total mass at t = 0 is conserved for t > 0. In fact, we have:
Z
¡1
1
u(x; t) dx =
Z
¡1
1
e
¡ct
1 + x
2
e
¡2ct
dx
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
z=xe
¡ct
Z
¡1
1
dz
1 + z
2
= π:
This is the same as the total mass at t = 0. Therefore, we can write:
Z
¡1
1
u(x; t) dx =
Z
¡1
1
u(x; 0) dx:
Note that since V (x) = cx, t he particles move with velocity proportional to x. Consequently, the
initial profile becomes fatter in time as the total mass remains the same.
-4 -2 0 2 4
0
0.2
0.4
0.6
0.8
1
For negative constant c, c < 0, the solution has the same structure. However, u(x; t) becomes
mor e concentrated a round x = 0 in this case. It is seen that for t !1, u(x; t) approaches the Dirac
delta function.
-4 -2 0 2 4
0
0.5
1
1.5
2
2.5
3
If α = 0, the equation is no longer conservative, and the solution is:
u(x; t) =
1
1 + x
2
e
¡2ct
:
This this case, we have
Z
¡1
1
u(x; t) dx =
Z
¡1
1
dx
1 + x
2
e
¡2ct
= πe
ct
>
Z
¡1
1
u(x; 0) dx:
2.5 Li near and semi-linear transp ort equatio ns 29
Exercise 2.37. Consider the following equation
u
t
+ e
x
u
x
= 0;
for t 0. Draw th e characte ristic curves an d nd the general solution of the equatio n.
Exercise 2.38. Consider the following equation
8
<
:
u
t
+ xu
x
¡ yu
y
= αu
u(x; y; 0) =
(
1 x
2
+ y
2
1
0 otherwise
:
a) For what value of α the equation is con servative?
b) For th is α, nd u(x; y; t).
c) General ize the above results in R
n
8
>
>
>
>
<
>
>
>
>
:
u
t
+
P
j=1
2n
(¡1)
j
@
j
u = αu
u(x; y; 0) =
(
1
P
j=1
2n
x
j
2
1
0 otherwise
:
Problem 2.4. In the ow of a inco mpressible fluid in a long pipeline, the velocity depends to the area intersectio n;
se e the following.
A
1
A
2
v
1
v
2
Use the conservation of mass principle, and show that
v
1
v
2
=
A
2
A
1
:
Problem 2.5. Suppose th e density function of a fluid is given by u(t; x; y) = e
¡t
(x
2
+ y
2
). Find the total mass
in the unit ball B = f(x; y); x
2
+ y
2
< 1g. Wha t is the rate of change of the total mass in B at time t = 1?
Problem 2.6. Suppose the initia l density function for a 1D ow is u
0
(x) = 1 + x. If the velocity eld for the
flow is v = t + x, we know that u(x; t) = e
¡2t
(1 + t + x) for t > 0. Verify the conservation law for this flow in an
ar bitrary segment [x
1
; x
2
].
Problem 2.7. For a 2D flow, suppose V = (x; y). If the init ial density fun ction is u
0
(x; y) = x
2
+ y
2
, then the
density function at t > 0 is u(t; x; y) = (x
2
+ y
2
) e
¡4t
. Verify directly the conservati on law for thi s flow.
Problem 2.8. Assume tha t the initial density of a matter is given by u
0
(x; y) = e
¡(x
2
+y
2
)
kg/m
2
. If the mass
moves w ith the constant speed V = 1m / s in the direction of x-axis, nd the total mass in the unite disk
(x ¡1)
2
+ y
2
< 1 the time t = 1.
Problem 2.9. Verify that the functi on u(x; y; t) = xy is a solution to the problem
(
u
t
+ xu
x
¡ yu
y
= 0
u(x; y; 0) = xy
:
Verify the conservation law
d
dt
Z
B
u(x; y; t) dS =
Z
bnd(B)
J
~
·νd`;
where B is the unit bal l in R
2
.
Problem 2.10. Write down the co ntinuity equation for function u for the velocit ies given below, and determine
if they are incompressible
a) V = ¡x
b) V = u
30 Li near First-Order PDEs
c) V = (y; x)
d) V = (1; 0; u)
Problem 2.11. Consider the veloci ty field V = (¡y; x) in R
2
. If the initial density is u
0
(x; y) =
1
1 + x
2
+y
2
, nd
the density of the point (¡3; 0) at t ime t = π without solving the tran sport equation. Find the total mass in the
unit disk at time t = 1.
Problem 2.12. Suppose the ini tial density fu nction of a 2D uid is given by u
0
(x; y) = e
¡x
2
¡y
2
. If th e fluid moves
with the velocity V = (¡y; 10x ¡2y), nd the parametric equations of a particle initially loca ted at (x
0
; y
0
) =(0; 3).
Find the de nsity of the point (x; y) = (0; 3e
¡π/3
) at time t =
π
3
.
Problem 2.13. Su ppose the velocity l ed of a 2D fluid is given by V = (1 + x
2
; 1). Find the ow line passi ng
through the origin and show that the density is decreasing along this line.
Problem 2.1 4. Write the continuity equation for a uid movi ng with the velocity V = (xy; y ¡x). Is the flow
incompressible? Repeat the problem for the vel oc ity le d V = (αx + y; e
x
¡ y) and nd α such th at the ow is
incompressible.
Problem 2.15. Assume that the density function of a fluid (in absence of any source term) is u(x; y; t) = e
¡t
(x
2
+
y
2
). If we know t he divergence of the velocity field V = (v
1
; v
2
) i s equal to 1, show the r elation xv
1
+ yv
2
= 0.
Problem 2.16. Assume that the initial density of a matter is given by u
0
(x) =
2
π(1 + x
2
)
kg/m. If the mass moves
with constant velocity V = 1m/s along x-a xis, nd the total mass in th e segment [0; 2] at t = 1.
Problem 2.1 7. Assume that the velocity of some fluid is gi ve n by V = x. Write down the continuity equation
for the flow and obtain the density at t = 1 if the initial density is u
0
(x) =
e
¡2
1 + x
2
.
Problem 2.18. Dete rmine the characte ris tic line along which th e solution to the problem
(
u
t
+ x u
x
= 0
u(x; 0) = e
¡jxj
;
is u = e
¡1
.
Problem 2.19. Draw the characteris tic lines of the equation
u
t
¡x u
x
= 0:
If the initial data is u(x; 0) = tan
¡1
x +
π
2
, find u(x; t).
Problem 2.20. Consider the equation
u
t
+ v(x)u
x
= 0;
where v(x) is the following function
v(x) =
8
>
>
<
>
>
:
¡x + 1 0 < x 1
¡x ¡1 ¡1 < x < 0
0 jxj1
:
If the initial data f is the function
f(x) =
jxj jxj1
1 jxj1
:
Problem 2.21. So lve the fo llowing linear probl em
(
u
t
+ (1 + x
2
) u
x
= t
u(x; 0) = x
:
Problem 2.22. Find the solution to the following
u
t
+ xu
x
= u
u(x; 0) = sin(x)
:
Problem 2.23. So lve the fo llowing equation
(
u
t
+ xu
x
= xe
¡t
u
u(x; 0) = sin(x)
:
2.5 Li near and semi-linear transp ort equatio ns 31
Problem 2.24. Consider the equation
3t
2/3
u
t
+ u
x
= 0;
for the function u = u(x; t).
a) Write th e e quati on of the characteris tic line passi ng through the point t = 1; x = 2.
b) Find the value u(1; 2) ¡u(4; 3) (you should provide the rea son for the value)
c) Find u(1; 2) if u(x; 0) = e
¡x
2
Problem 2.25. Consider the foll owing se mi-linear problem
(
u
t
+ u
x
= 2te
¡u
u(x; 0) = ln(1 + x
2
)
:
Find region in the plane (x; t) where u < 1.
Problem 2.26. Assume that the velocity el d of a matter is give as V = (¡y; x). Find the density of the point
(1; 0) at time t = 1 if the plan (x; y) is initially lle d with a matter of density
i. u
0
= x
2
+ y
2
p
,
ii. u
0
= xy
Problem 2.27. So lve the fo llowing problem
(
u
t
+ u
x
+ yu
y
= xy
u(x; y; 0) = sin(xy)
:
Problem 2.28. So lve the fo llowing problem
(
u
t
+ u
x
+ u
y
= u
u(x; y; 0) = u
0
(x; y)
:
Problem 2.29. Consider the equation
u
t
+ yu
x
¡xu
y
= 0:
a) Show that the equa tions of characteristic curves are
x = x
0
cos(t) + y
0
sin(t)
y = ¡x
0
sin(t) + y
0
cos(t)
:
b) Find the solution if u(x; y; 0) = xy.
c) Now solve the following equation
(
u
t
+ yu
x
¡xu
y
= u
u(x; y; 0) = xy
:
2.6 Qua si- linear equations and shockwave
In general, V in the continuity equation u
t
+ r·[uV ] = 0 can be a function of x and u, such that
V = V (x; u). One example of this is in modeling traffic flow on highways, where the velocity of cars
depends inversely on the density of cars u sing the h ighway, describe d by the function
V (u) = V
max
1 ¡
u
u
1
;
where u
1
is the density of c ars at maximum capacity, V
max
is the maximum speed of cars, and
0 u u
1
repre sents the density of traffic on the highway. In this case, the continuity equation
becomes
u
t
+ V
max
1 ¡
2u
u
1
u
x
= 0: (2.22)
32 Li near First-Order PDEs
Anot her important equation with V as a function of u is the B urgers' equation, where V =
1
2
u,
giving the PDE
u
t
+
1
2
u
2
x
= 0;
or equivalently, u
t
+ uu
x
= 0.
More generally, we can consider problems of the form:
u
t
+div [g(u)] = 0
u(x; 0) = u
0
(x)
;
which can be written equivalently as:
u
t
+ g
0
(u) u
x
= 0
u(x; 0) = u
0
(x)
:
The characteristic system for this problem is given by:
8
<
:
dx
dt
= g
0
(u)
du
dt
= 0
:
which leads t o the following implicit solution: u = u
0
(x ¡ g(u)t).
Example 2.25. Let's consider the continuity equation problem given by
u
t
+uu
x
=0;
with the initial condition u(0; x)=x. Although the initial condition u(0; x) for x<0 is not physically
meaningful a s i t leads to negative density, we retain this example to highlight an important feature
of the continuity equation. By solving the characteristic system of equations
dx
dt
= u;
du
dt
= 0;
we obtain the solution x=ut+x
0
and u(t; x(t))=u(0; x(0))=x
0
, where x
0
is replaced by x ¡ut.
This yields the solution
u(x; t)=
x
1 + t
;
which has a domain of [0; 1) for t>0.
Suppose we now modify the init ial condition to ¡x. The resul ting solution is
u(x; t)=
x
t ¡1
;
which has a domain of [0; 1). As shown in the figure, the characteristic lines in this case collid e at
t=1, leading to a shock in the solution.
-1 -0.5 0 0.5 1
0
0.2
0.4
0.6
0.8
1
2.6 Quasi-linear equations and shoc kwave 33
The value of u(1; 0) can take any value in the range (¡1; 1), as particles with different densities
converge at x=0 at t=1 and the density becomes discontinuous at this point. The figure below
shows this phenomena schematically. This concentration changes the initial densi ty of x=0 to a
discontinuous value.
x = 0
x
u(1; x)
t = 1 u
0
(x
0
)
x
0
t = 1
u
0
(¡x
0
)
¡x
0
Exercise 2.39. Sol ve the following quasi-linear equation and determine the domain of the solution
u
t
+ uu
x
= 0
u(x; 0) = jxj
:
Remark 2.26. What can we say about the solution u(x; t) for t > 1? As we have seen, the solution
is discontinuous at t = 1, which means that the differential equation fails to hold for t 1 as well.
However, this does not mean that we cannot study the propagation of the shock beyond t = 1 from
a physical point of view. In fact, the development of a shock is a physical p henomenon that can b e
observed and studied experimentally.
One possible mathematical way to extend the discontinuous or shock solutions beyond the shock
time t = 1 is by using the concept of shockwave solutions. In this approach, we assume that the
discontinuous solution propagates like a wave in time. Shockwave solutions are a common way of
extending the solutions of partial d ifferential equations beyond t he point of discontinu ity.
2.6.1 Riemann problem
In this section, we consider the initial value problem given by the scalar conservatio n law:
8
>
>
>
>
<
>
>
>
>
:
u
t
+ uu
x
= 0
u(x; 0) = u
0
(x): =
8
>
>
<
>
>
:
1 x < 0
1 ¡x 0 < x < 1
0 x > 1
: (2.23)
To solve this problem, we first determine the characteristic system of equations:
8
<
:
dx
dt
= u
du
dt
= 0
;
which leads to the characteristic lines x(t)=ut+x
0
. Along each charact eristic line, the solution u
remains constant. Thus, we can write:
u(t; x(t)) = u
0
(x
0
);
where x
0
is the initial position of the characteristic line. Using the initial condition u
0
(x), we can
find the value of u along e ach characteristic line.
34 Li near First-Order PDEs
Next, we express x
0
in terms of x and t as x
0
=x¡ut. Substituting this into the expression for
u(t; x(t)), we obtain the solution u as a function of x; u and t:
u(x; t) =
8
<
:
1 x ¡ut < 0
1 ¡(x ¡ut) 0 < x ¡ut < 1
0 x ¡ut > 1
:
Simplifying the expression algebraically, we get:
u(x; t) =
8
>
>
>
>
<
>
>
>
>
:
1 x < t
1 ¡x
1 ¡t
t < x < 1
0 x > 1
: (2.24)
The figure below shows the solution u(x; t) at several instances of time. Note how a shock (discon-
tinuity) is developed at t=1.
t = 0
u
x
t = 0.5
u
x
x
u
1
x
t = 0.75
1
u
1
1
t = 1
discontinuity at t=1
Let's analyze the Riemann problem again using equation (2.23). In this equation, the function u
has a dual role, serving both as the density function and the material velocity V , where V =
dx
dt
= u.
Now, let's examine the initial condition shown below at t = 0:
t = 0
u
1
x
Particles initially located at x
0
0 move to the right with velocity V =1, as their density
u =1 is conserved. For example, a p article initially located at x
0
=0 will reach x=1 at t = 1,
as evident from the characteristic equation x(t)=ut+x
0
. When x
0
= 0 and u=1, x(1)=1.
Particles located initially at x 2(0; 1) move with the veloc ity V = 1 ¡x while they preserving
their densities. For example, a particle at x
0
=
1
2
has density u=
1
2
and moves with speed
V =
1
2
to the right, reaching x=0 at t=1.
All particles initially located at x 1 have zero velocity and remain stationary on the x-axis
for t >0.
2.6 Quasi-linear equations and shoc kwave 35
The above observations imply that the density of point x=1 at time t = 1 can take any value between
0 and 1, as particles with different densities between 0 and 1 reach this point at t=1. Therefore,
we can write the density u(x; 1) as
u(x; 1) =
1 x < 1
0 x > 1
;
while u(1; 1) is not uniquely defined.
u
1
x
t = 1
2.6.2 Extension of the solution
We will delve into the details of this problem in the second volume of the book. However, in this
context, we aim to understand the elements of shock propag ation through the conservation law. Our
primary concern is to determine the density function u(x; t) for t>1. It should be noted that the
partial differential equation is not valid at t = 1, and hence, we cannot use the solution (2.24) for
t > 1. In fact, using this solution for t > 1 results in the existence of multiple non-physical solutions,
as shown in the figure below.
t = 2
u
x
1
1
x
u
t = 1.5
To unde rstand the p ropagation of shocks better, let us first consider the following one-dimen-
siona l wave equation:
u
t
+ cu
x
= 0
u(x; 0) = u
0
(x)
;
where u
0
is the following discontinuous function
u
0
(x) =
1 x < 0
0 x > 0
:
As expected, the wave equation carries the initial condition with speed c to the right or left depending
on the sign of c in t he equation. Hence, the solution can be expressed as:
u(x; t) = u
0
(x ¡ct):
The following figure shows the solution of the wave equation for t = 1 and t = 2. Note that the
extension of the initial condition is plausible, and the solution is discontinuous only at one point
for each time instance.
t = 1
u
t = 2
u
1 + c1
x
x
36 Li near First-Order PDEs
Let's return to the Riemann problem:
8
<
:
u
t
+ uu
x
= 0
u(x; 1) =
1 x < 1
0 x > 1
:
Since u can only take the values 0 or 1, except at the discontinuity point, it's reasonable to assume
that this equ ation behaves like a wave equation. If we make this assumption, then the question
becomes: at what speed will the initial condition propagate?
To answer this, we can use the conservation law and write the PDE as an integral equation:
Z
x
0
x
1
u
t
+
1
2
u
2
x
dx = 0:
for any interval (x
0
; x
1
) containing the shock position x at time t. Integrating this equation over
the interval (x ¡"; x + "), where " is a small positive constant, we obtain
Z
x¡"
x+"
u
t
(x; t) dx +
1
2
u
r
2
¡
1
2
u
l
2
= 0; (2.25)
where u
l
and u
r
are the values of u to the left and right of the shock, respectively. On the other
hand, we can write the integral in the above equation as
Z
x¡"
x+"
u
t
(x; t) dx = lim
δt!0
1
δt
Z
x¡"
x+"
u(x; t + δt) dx ¡
Z
x¡"
x+"
u(x; t) dx
:
The first integral in the right-hand side of the above equation is equal to
Z
x¡"
x+"
u(x; t + δt) dx = u
l
(" + δx) + u
r
(" ¡δx);
and the second one is
Z
x¡"
x+"
u(x; t) dx = u
l
" + u
r
":
The scenario is shown in the figure below.
t
x + δx
x
x
x ¡" x + "
x + "
x ¡"
t + δt
Hence, we can write
Z
x¡"
x+"
u
t
(x; t) dx = v(u
l
+ u
r
);
where v is the ve locity of the shock point x. Substituting the result into equation (2.25) yields
v =
1
2
v
l
2
¡
1
2
u
r
2
u
l
+ u
r
=
u
l
+ u
r
2
:
For our example with u
l
= 1 and u
r
= 0, we obtain the shock velocity v =
1
2
. The solution for t = 1;
2 is illustrated in the following figure:
2.6 Quasi-linear equations and shoc kwave 37
t = 1
u
t = 2
u
1
x
3
2
x
The solution u(x; t) can be obtained using the shockwave solution of the Riemann problem. The
formula for the solutio n is given as:
u(x; t) =
1 x < 1 + v(t ¡1)
0 x > 1 + v(t ¡1)
;
where v =
1
2
is the shock velocity for the g iven example.
This solution is called the shockwave solution be cause it has a sharp transition between the two
states, which moves with constant velocity v. The following figure illustrates the characteristic lines
with the shock l ine.
x
= 1
u = 1
x
t
= 1
u = 0
t
t = 2x ¡1
Example 2.27. This example illustrate the fact that t he graph of shockwave in xt-plane can be
a curve. This means that the ve locity of shock propagation can vary with time. Let us solve the
following problem
8
<
:
u
t
+ uu
x
= 0
u(x; 0) =
1 + x ¡1 < x < 0
1 ¡x 0 < x < 1
:
The equation of characteristic lines are x(t) = u
0
(x
0
) t + x
0
where u
0
(x
0
) =
1 + x
0
¡1 < x
0
< 0
1 ¡x
0
0 < x
0
< 1
. Sub-
stituting u
0
(x
0
) into the equation of characteristic line yields
x(t) = x
0
+
(1 + x
0
)t ¡1 < x
0
< 0
(1 ¡x
0
)t 0 < x
0
< 1
:
A simple algebraic calculation helps to solve x
0
in terms of x; t as follows
x
0
=
8
>
>
<
>
>
:
x ¡t
1 + t
¡1 < x < t
x ¡t
1 ¡t
t < x < 1
:
Accordingly, the solution u(x; t) is obtained as
u(x; t) =
8
>
>
<
>
>
:
1 + x
1 + t
¡1 < x < t
1 ¡x
1 ¡t
t < x < 1
:
38 Li near First-Order PDEs
The solution for t =
1
2
;
3
4
and t = 1 are shown below.
1
.
0 0
.
5 1
.
0
1
t
= 0
.
5
1
.
0
0
.
75
1
t
= 0
.
75
1 1
1
t
= 1
t is observed that at t = 1, a discontinuity is formed at x
= 1. For t > 1, the solution extends as
a sho ckwave with speed v that needs to be determined. Similar to the argument above, it can be
justified that the shockwave propagates with speed v =
1
2
at t = 1. However, what about v for t > 1?
It is important to note that according to the conservation law, the total mass of the system
remains conserved. This can be expressed as:
Z
¡1
1
u(x; t) dx =
Z
¡1
1
u(x; 0) dx = 1:
If the sho ckwave were to move with a constant velocity v =
1
2
for all t > 1, then the area under the
curve u(x; t) would increase with time, which would violate the conservation of mass. In order to
maintain the constant mass equal to 1, the height of the triangle, h, must decrease with time. This
is depicted in the following figure.
1
u
¡1 1.45
x
2.74
t = 2
t = 1
t = 6
h = 2/ 14
p
h = 1
h = 2/ 6
p
It is important to note that the velocity of the shockwave is not constant in the case where the
initial data has a jump discont inuity. Since the left value of the solution is h and the right value is
0, we can derive the shock velocity as v =
1
2
h, which in turn gives v =
1
1 + x
, where x is the distance
from the discontinuity at x = 1. Therefore, the shock velo city varies with respect to x.
To find the shock velocity in terms of t, we can integrate the relation v =
dx
dt
=
1
1 + x
with resp ect
to t. This gives us x(t) = 2t + 2
p
¡1, which implies that the shock velocity at time t is v =
1
2t + 2
p
for t 1. The characteristic li nes in the (x; t)-plane can be plotted to visualize the behavior of the
solution.
1 0 1 2
1
2
3
shock
2.6 Quasi-linear equations and shoc kwave 39
Exercise 2.40. An alternative method to nd the shockwave velocity for the above problem is to use the following
formula
u
l
(x; t) =
x + 1
t + 1
;
and to solve the equa tion
v =
dx
dt
=
x + 1
2(t + 1)
:
Carry out the calculati on and conclude v =
1
2t + 2
p
.
Exercise 2.41. Let us solve the foll owing problem
u
t
+ uu
x
= 0
u(x; 0) = u
0
(x)
;
where u
0
is as follows
u
0
(x) =
3 x < 0
1 x > 0
:
Draw the ch ar ac teristic lines in the (x; t)-plan e and find the shockwave solution.
Exercise 2.42. Consider the following problem
u
t
+ uu
x
= 0
u(x; 0) = u
0
(x)
;
where u
0
is a continuously differentiable fun ction.
a) Show that if u
0
< 0, the equation will develop a shock.
b) show that the rst time of the appearance of the shock is derive d by the following formula
t
=
¡1
min
ξ
u
0
0
(ξ)
Hint : The shock is formed i f two character istic li nes collide
x
0
+ u
0
(x
0
)t = x
1
+ u
0
(x
1
)t:
This gives t =
x
0
¡ x
1
u
0
(x
1
) ¡ u
0
(x
0
)
. Minimize this and conclude the above rela tion.
Exercise 2.43. Consider the following equation
u
t
+ g
0
(u) u
x
= 0
u(x; 0) = u
0
(x)
:
If the probl em devel ops a shock, show that the velocity of the shockwave is
v =
g(u
l
) ¡ g(u
r
)
u
l
+ u
r
;
where u
l
; u
r
ar e the left and right limit of u at the shock point x respectively.
Exercise 2.44. Consider the following problem
8
>
>
>
>
<
>
>
>
>
:
u
t
+ uu
x
= 0
u(x; 0) =
8
>
>
<
>
>
:
0 x 0
x
"
0 x "
1 x "
;
for " > 0. The problem does not devel op shock.
a) Draw the initial condition and obtain the solution u
"
(x; t).
b) Let " !0 and find the solution of the follow ing equation
8
>
>
>
>
<
>
>
>
>
:
u
t
+ uu
x
= 0
u(x; 0) =
8
>
>
<
>
>
:
0 x < 0
1 x > 0
:
The solution of this equation is called the rarefaction solution.
40 Li near First-Order PDEs
Problem 2.30. Find an explicit solution of the followi ng equation
u
t
+ uu
x
= x
u(x; 0) = 0
:
Problem 2.31. Find the explicit solution to the followin g problem
(
u
t
+ e
u¡x
u
x
= 0
u(x; 0) = x
:
Problem 2.32. So lve the fo llowing problem
u
t
+ uu
x
= u
u(x; 0) = 1 + x
:
Problem 2.33. Find an implicit soluti on to the following problem
u
t
+ xuu
x
= 0
u(x; 0) = x
:
Use the implicit function theorem to justify that the solution exists in some interval of t 2[0; T ).
Problem 2.34. Find an implicit soluti on to the following problem
@
t
u + sin(u) @
x
u = 1
u(x; 0) = x
:
Use the implicit function theorem to justify that the solution exists in some interval of t.
Problem 2.35. Draw the solution to the following problems for t = 0; 1; 2
u + uu
x
= 0
u(x; 0) = u
0
(x)
;
where
u
0
(x) =
8
>
>
<
>
>
:
0 x 0
x 0 x 1
1 x 1
:
Problem 2.36. Consider the foll owing problem
8
>
>
>
>
<
>
>
>
>
:
u
t
+ uu
x
= 0
u(x; 0) =
8
>
>
<
>
>
:
2 x 0
2 ¡x 0 x 1
1 x 1
:
i. Draw t he characteri stic lines and determine t
.
ii. Find the shock wave speed.
iii. Draw the solution u(x; 0 .5), u(x; 1), u(x; 2) and u(x; 4).
Problem 2.37. Draw the solution u(1; x) and u(2; x) to the following problem
8
<
:
u + uu
x
= 0
u(x; 0) =
2 x 0
1 x 0
:
Problem 2.38. So lve the fo llowing problem for t < t
and draw the solutio n for few values o f t.
8
<
:
u
t
+ uu
x
= 0
u(x; 0) =
(
1 + x ¡1 < x < 0
e
¡x
0 x
:
Problem 2.39. For th e following equa tion, draw the charac teristics and nd the shock wave speed. Draw the
sol ut ion u(1; x) and u(2; x):
8
<
:
u
t
+ uu
x
= 0
u(x; 0) =
2 ¡jxj ¡1 < x 1
0 otherwise
:
2.6 Quasi-linear equations and shoc kwave 41
Problem 2.40. The traffic ow in highways is usually modeled by the fol lowing equation
u
t
+ (1 ¡u)u
x
= 0;
where u is the density of cars (number of cars in a unit length). Assume that the initial profile is as follows
u(x; 0) =
8
>
>
<
>
>
:
1 x 0
x + 1 0 x 1
2 x 1
:
a) Draw the characteristic lines of the eq uation and find the collision time t
?
.
b) Dr aw u(x; t) where T = 2.
Problem 2.41. For the fol lowing equation, nd the collision t ime t
. Use a software to ne u(x; 0.5) and u(x; 1).
(
u
t
+ 2uu
x
= 0
u(x; 0) = e
¡jxj
:
Problem 2.42. So lve the fo llowing problem
8
<
:
u
t
+ uu
x
= 0
u(x; 0) =
1 x < 0
2 x > 0
:
Problem 2.43. So lve the fo llowing problem
8
>
>
>
>
<
>
>
>
>
:
u
t
+ uu
x
= 0
u(x; 0) =
8
>
>
<
>
>
:
0 x < 0
1 0 < x < 1
0 x > 1
:
42 Li near First-Order PDEs
Appendix A
C urves and Surface
A.1 Curves and surfaces
The techniques outlined in this chapter for solving first-order PDEs are primarily based on the
concept of the derivative of a scalar function along a smooth curve. To apply these techniques, it is
necessary to have an understanding of how smooth curves and surfaces are represente d in the plane
or in R
n
for n > 2.
A.1.1 Curves and derivative along a curve
Curves in the plane R
2
can be represented in several ways: through an explicit funct ion as the graph
f(x; f(x)); x 2D
f
g where f is a smooth function, an implicit func tion φ(x; y)=c, or a parametric
form γ(t)=(x(t); y(t)), where t ranges over an open interval I R. An implicit representation φ(x;
y)=c o f a curve can be transformed into an explicit one, y = f (x), under certain conditions using
the implicit function theorem; see the appendix of the book. Note that the graph of an explicit
function y = f(x) is a special case of a parameterized curve where the parameter is chosen to be x,
so that γ(x)=(x; f(x)).
While explicit and implicit representations of curves focus on the curves as a geometrical object,
or a set of points in the xy-plane, parametrized c urves can be considered as the path or trajectory
of a moving particle in the plane. Accordingly, the parameter t represents time, and the function
γ(t) gives the position of the particle at each time t. This interpreta tion enables us to determine the
veloc ity and acceleration vectors of the particle along its path. Moreover, the concept of parame-
trization can be generalized to higher dimensions, allowing us to represent more complicate d curves
and surfaces in R
n
.
A curve in R
n
parametrized by a parameter t is define d as a map γ: I R !R
n
given by
γ(t)=(x(t); x(t); :::; x(t)). Here, x; x; :::; x are scalar functions of t that describe the coordinate s
of points on the curve, and t is a parameter that varies over an open interval I R. The tangent
vector at a specific point γ(t) is defined as the derivative of γ(t) with respect to t, written as
γ
0
(t)=(x
1
0
(t); x
2
0
(t); :::; x
n
0
(t)). The magnitude of this vector is d enoted by jγ
0
(t)j, and in the c ontext
of physics, it repr esents the speed of a moving particle along γ(t).
The arc length s( t) of a curve is defined as the integral of the speed jγ
0
(t)j over the interval [0; t],
that is,
s(t)=
Z
0
t
jγ
0
(τ)j :
By the fundamental theorem of calculus, the derivative of s(t) with respect to t is given by
ds
dt
= jγ
0
(t)j;
43
which is the speed of a moving particle along the curve γ(t).
The equation of a straight line γ = (x
1
(t); : : : ; x
n
(t)) in R
n
passing through c = (c
1
; : : : ; c
n
) is
simply derived by the differential equations
dt
= a;
where a = (a
1
; : : : ; a
n
) is a constant. In the coordinate system, we have
dx
j
dt
= a
j
for j = 1; : : : ; n,
which is solve for x
j
= a
j
t + c
j
, or in the algebrai c form:
x
1
¡c
1
a
1
= ··· =
x
n
¡c
n
a
n
:
The vector a determines the direction of the line and is tange nt to γ as γ
0
(t) = a.
For a general curve γ(t) whose direction changes continuously according to the vector function
V (x) = (v
1
(x); : : : ; v
n
(x)), where x =(x
1
; :::; x
n
), the equation is derived by solving the following
differential equati on:
dγ
dt
= V (γ(t));
or in the coordinate system,
dx
j
dt
= V (x
1
(t); : : : ; x
n
(t)); for j =1; :::; n, where γ(t)=(x
1
(t); :::; x
n
(t)).
If γ(t) is considered as the trajectory of a particle in R
n
, the vector V (γ(t)) is the velocity of the
particle at the point γ(t), and the vector function V = (v
1
;:::; v
n
) is said to be the velocity vector field.
Example A.1. Consider a partic le moving in the xy-plane with the directional vector function V (x;
y)=(¡y; x). The trajectory γ(t) of the particle can be derived by solving the system of differential
equations:
dx
dt
= ¡y;
dy
dx
= x:
This system is equivalent to the exact equation xdx + ydy = 0 in the xy-plane, which has the implicit
solution x
2
+ y
2
= c. Alternatively, the system can be converted to a second-order equation for x(t)
as x
00
+ x = 0, with solution
x(t) = x
0
cos(t) ¡ y
0
sin(t);
where x
0
is the initial position of the particle (i.e., x(0)=x
0
), and y
0
= ¡x
0
(0). Using x(t) and the
second differential equation, we can find
y(t) = x
0
sin(t) + y
0
cos(t);
and thus the equation of the trajectory is obtained as
γ(t) = (x
0
cos(t) ¡y
0
sin(t); x
0
sin(t) + y
0
cos(t)):
In particular, if x
0
= 1 and y
0
= 0, we obtain the trajectory
γ
(1;0)
(t) = (cos(t); ¡sin(t)):
As we can see, the trajectory of the particle is the unit circle in the xy-plane. We can alternatively
solve the system by putting it in the matrix form
d
dt
x
y
=
0 ¡1
1 0

x
y
;
44 Curves and Surface
with the fu ndamental matrix Φ(t) given by:
Φ(t) =
cos(t) ¡sin(t)
sin(t) cos(t)
:
Note that:
Φ(t)
1
0
= γ
(1;0)
(t):
The velocity vector of the particle with the trajectory γ
(1;0)
(t) is
γ
(1;0)
0
(t) = (¡sin(t); cos(t));
and throu gh the relation
γ
(1;0)
(t) · γ
(1;0)
0
(t) = 0;
we observe that γ
(1;0)
0
(t) is perpendicular to the traj ectory of the particle.
Let γ(t) be a parametric curve in R
n
, and let u = u(x
1
; : : : ; x
n
) be a continuously differentiable
function. The derivative of u along γ(t) is defined as:
du
dt
:=
X
j=1
n
@
j
u(γ(t))
dx
j
dt
= ru(γ(t)) · γ
0
(t);
where ru denotes the gradient of u, given by the vector:
ru =
0
B
B
@
@
1
u
·
·
·
@
n
u
1
C
C
A
:
The operator r is also known as nabla and plays an important role in the context of multivariable
functions. Fo r example, if u represents the density f unction in R
n
, and γ(t) is a smooth curve, then
du
dt
(γ(t)) represents the rate of change of the density when a control volume of mass is moving along
γ.
Example A.2. The above explanation highlights the relationship between curves in R
n
and systems
of first-order ordinary differential equations. Specifically, a curve in R
n
can be obtained by solving
a system of first-order ordinary differential equations, and conversely, the solution of a system of
ordinary differential equations corresponds to a curve. This c onnection between curves and differ-
ential equations is fundamental to many areas of mathematics and has wide-ranging applications
in physics, engineering, and other fields.
Consider the system
8
<
:
dx
dt
= ¡y ¡αx
dy
dt
= x ¡αy
;
where α is a constant. The trajectory associated with α = 0; 0.3; ¡0.3, are shown b elow
A.1 Curv es and surfaces 45
In the xy-plane, the equation the equation can be written as
(x ¡αy) dx + (y + αx)dy = 0;
which is not exact if α =/ 0, meaning that energy is not conserved along the paths of mo tion. The
equivalent se cond-order equation of the system is:
x
00
+ 2αx
0
+ x = 0;
which can which can be further transformed to:
d
dt
1
2
jx
0
j
2
+
1
2
x
2
= ¡2α jx
0
j
2
:
Integrating this equation yields:
E(t) = E(0) ¡2 α
Z
0
t
jx
0
(s)j
2
ds:
If α > 0, the particles move down spirally to the origin along V (x; y), while if α < 0, they gain energy
and move outward spirally.
Exercise A.1. Find the force eld associated to th e velocity led given in the example above using t he relation
F =mγ
00
(t), and show that this force field is centrifugal. Consider the t rajectory sta rting at (1; 0) for s implicity.
Additionally, demonstrate that the energy of the particle is conserved along its trajectory for the energy given by
E(t) = mjγ(t)j
2
+
1
2
mjγ
0
(t)j
2
:
Exercise A.2. Co nsider the vector eld V = (¡y; x; 1) and suppose a par ticle located initially at (1; 0; 0) moves
according to this vector eld. Find the trajectory of the particle an d show the relation γ
0
(t) · γ
00
(t) = 0 for all t.
Thus, conclude that the frame
γ
0
(t)
jγ
0
(t)j
;
γ
00
(t)
jγ
00
(t)j
;
γ
0
(t) × γ
00
(t)
jγ
0
(t) × γ
00
(t)j
defines a local coordinate system for the particle.
46 Curves and Surface
Exercise A.3. With a curve mapping, we can determine the length of a curve. For a given explicit curve y = f (x)
or (x; f(x)) for x 2(x
0
; x
1
) the differential length is
ds = 1 + jf
0
j
2
p
dx;
and thus the arc length i s
s =
Z
x
0
x
1
1 + jf
0
j
2
p
dx:
Simi larly, for a general parametriza tion γ(t), the differential le ngth is
ds = jγ
0
(t)jdt = jx
0
(t)j
2
+ jy
0
(t)j
2
p
dt;
and thus
s =
Z
t
0
t
1
jx
0
(t)j
2
+ jy
0
(t)j
2
p
dt:
The interesting point is that the arc l ength formula is independent of the parametriza tion. Let C R
2
be a
geometric curve and let γ
1
: (a; b) !C and γ
2
: (c; d) !C be two smooth parametrization of C. Show
Z
a
b
jγ
1
0
(t)jdt =
Z
c
d
jγ
2
0
(t)jdt:
Exercise A.4. Consider the temperature dist ribution in the xy-plane given by T (x; y)=10 exp(x
2
¡y
2
). Suppose
a runner is moving along the curve γ(t)=(cos(t); sin(t)), and wearing a hand clock m arke d from 0 to 60. What is
the angular velocity of the hand of the runner's watch as they move alon g the curve? Find the angular ve locity
of the hand clock if the runn er runs given by the following function γ(t) = (cos(!
0
t); sin(!
0
t)) for some !
0
> 0.
A.1.2 Surfaces in space
There are several ways to represent a surface in three-dimensional space. One common approach
is to express the surface as the graph of an explicit function of the form z = f(x; y), where z is the
dependent variable and x; y are the independent variables. Alternatively, the surface can be defined
implicitly through an equation of the form φ(x; y; z)=c.
The first approach is a convenient way to visualize and manipulate the surface. However, this
representation is only possible for surfaces that can be written as a function of x and y, such as
paraboloids or planes. The second approach is more general and can represent surfaces that cannot
be written explicitly as a function of x and y, such as spheres or tori. The implicit function theorem
provides conditions for the existence and differentiability of a smooth function z = f(x; y) that
satisfies the equation φ(x; y; f (x; y)) = c. In particular, the the orem requires the existence of a
point (x
0
; y
0
; z
0
) such that φ(x
0
; y
0
; z
0
) = c and the partial derivative
@φ
@z
(x
0
; y
0
; z
0
) is nonzero. This
condition guarantees that the equation φ(x; y; z)=c defines the surface locally as a graph over the
xy-plane, and the function f (x; y) can be obtained by solving for z in terms of x and y.
Finally, a surface can be parameterized using a set of e quations that describe how x; y, a nd z
vary with two independent variables, typically denoted by t and s. In particular, a parameterized
surface in three-dimensional space is given by the equation
Σ(t; s)=(x(t; s); y(t; s); z(t; s));
where x; y, and z are functions o f the independent variables t and s, and (t; s) belongs to an open
set D in the plane. This method allows for the representation of surfaces with complex shapes
and topologies, such as the torus or the Mobius strip, and can also be used to represent higher-
dimensional surfaces. For instance, an m-dimensional hypersurface in R
n
where m < n can be
represented by a vector function of the f orm
Σ(t
1
; : : : ; t
m
) = (x
1
(t
1
; : : : ; t
m
); : : : ; x
n
(t
1
; : : : ; t
m
)):
A.1 Curv es and surfaces 47
Moreover, the parametric representation of a surface reveals the relationship between curves and
surfaces. For example, a surface Σ( t; s) in R
3
contains the lines γ
s
0
(t) = (x(t; s
0
); y(t; s
0
); z( t; s
0
)),
where t ranges over an interval for each s
0
, and the curves γ
t
0
(s) = (x(t
0
; s); y(t
0
; s); z(t
0
; s)),
where s ranges over an interval for each t
0
. For instance, let D =[0; π]×[¡π; π], and consider the
parametrization given by:
Σ(t; s)=(sint coss; sint sins; cost);
for (t; s) in D. This parametrization maps the rectangle D onto the surface of the unit sphere . Each
line fs
0
g×[0; π], for s2[¡π; π], is mappe d to the vertical curve Σ(t; s
0
), and each line [¡π; π]×ft
0
g
to the horizontal curves Σ(s; t
0
). In the spherical coordinate system, t is usually denoted by θ, the
angle with the z-axis, and s by φ, the ang le with the x-axis.
It is worth noting that the graph of an explicit fu nction z =f (x; y) is actually a special case of
a parameterized surface, where the surface is defined by the equation Σ(x; y)=(x; y; f(x; y)). The
tangent and normal vectors to a surface Σ can be found as follows. Let (t
0
; s
0
) be a fixed point on
Σ. The curve map γ
s
0
(t)=(x(t; s
0
); y(t; s
0
); z(t; s
0
)) lies on Σ, and its tangent vector γ
s
0
0
(t
0
) is equal
to Σ
t
(t
0
; s
0
). Similarly, the curve map γ
t
0
(s)=(x(t
0
; s); y(t
0
; s); z(t
0
; s)) lies on Σ, and its tangent
vector is Σ
s
(t
0
; s
0
). The space spanned by Σ
t
(t
0
; s
0
) and Σ
s
(t
0
; s
0
) is the tangent space to the surface
Σ at ( t
0
; s
0
), which is just the tangent plane to Σ at that point. A surface Σ is said to be smooth if
Σ
t
(t; s) and Σ
s
(t; s) exist and are linearl y independent for all (t; s) in the domain of Σ. The normal
vector to Σ is then given by the cross product
ν =
Σ
t
×Σ
s
jΣ
t
×Σ
s
j
:
Note that if Σ
t
(t
0
; s
0
) and Σ
s
(t
0
; s
0
) are linearly independent, then Σ
t
(t
0
; s
0
) ×Σ
s
(t
0
; s
0
)=/ 0, and
the normal vector is well-defined. For example, the normal vector to the unit sphere at the point
Σ(π /2; 0)=(1; 0; 0) with the parametrizati on given above can be calculated as follows: Σ
t
(π /2;
0)=(0; 0; ¡1) and Σ
s
(π /2; 0)=(0; 1; 0). Therefore, Σ
t
(π /2; 0)×Σ
s
(π /2; 0)=(1; 0; 0) a nd jΣ
t
(π /2;
0)×Σ
s
(π/2; 0)j=1, so the normal vector is ν =
Σ
t
×Σ
s
jΣ
t
×Σ
s
j
= (1; 0; 0), which is clearly perpendicular to
the tangent plane of the sphere at the point Σ(π/2; 0).
Exercise A.5. Th e are a differential dS is equal to jΣ
t
×Σ
s
jdtds, and thus the area of the surface Σ is equal to
S =
ZZ
D
jΣ
t
×Σ
s
jdtds:
In particular, dS = 1 + jruj
2
p
dxdy is the differential area of an explicit sur face z = u(x; y). For example, if Σ
represent s a mat erial surface o f density ρ(x; y; z), th en the t otal mass on Σ is equal to
M =
ZZ
D
ρ(Σ(t; s))jΣ
t
×Σ
s
jdtds:
48 Curves and Surface
Use the spherical coordinate and and calculate the area of the pot ion 0 θ θ
0
of the unit sphe re for θ
0
<
π
2
.
Calc ulate the total mass on the surfac e if ρ = (x
2
+ y
2
)z.
Exercise A.6. Write down the equation of th e tangent plane to the smooth surface φ(x; y; z) = 0 at (x
0
; y
0
; z
0
)
on the surface.
Exercise A.7. Let φ(x; y; z) = 0 denote a surfac e Σ. Show that rφ is perpendicular to all tange nt vectors on
the surface Σ. Hint: Let γ = (x(t); y(t); z(t)) be arbitrary smooth curve lying on Σ .
Let's now consider the equations of surfaces in R
3
. Suppo se we have a plane passing through
a point p
0
= (x
0
; y
0
; z
0
) and perpendicular to a constant vector n~ = (a; b; c). Its equation can be
expressed as a
a(x ¡x
0
)+b(y ¡ y
0
)+c(z ¡z
0
)=0;
which is the expansion of the dot product (p ¡p
0
) ·n~ = 0, for p = (x; y; z) on the plane. Equivalently,
if we are given two linearly independent vectors V
1
= (a
1
; b
1
; c
1
) and V
2
= (a
2
; b
2
; c
2
), we can write
the parametric equation of the plane containing these two vectors as
Σ(t; s)=(ta
1
+sa
2
; tb
1
+sb
2
; tc
1
+sc
2
):
Note both vectors lie in the obtained plane. If we were provided with only one vector, the equation
of the plane could not be uniquely determined.
A.1 Curv es and surfaces 49