You are on page 1of 65

Communication & Computer Engineering Department

Subject: Advanced Engineering Mathematics


Dr. Bessam Z. Hassan

Chapter 3: Vector Calculus and Integration Theorems

3.1 Differentiation of Vectors


Previously position vectors were discussed in detail and we saw that a position vector gives the displacement from the origin as well as the x-, y-, and z- coordinates of a particle locating its position in the 3D space as shown in Fig. 3.1. Now if this particle is moving with respect to time then its x-, y- and z- coordinates are functions of time. Furthermore the particle will have velocity and might have acceleration as well which can both be presented by vectors.

Fig. 3.1 The Position vector of a particle moving through space is a function of time

In this section we will restrict the definition of the vector function to the domain of a single real variable t. This will result only in 3-D curves as shown in Fig. 3.2. Later on, we will define vector functions with a domain in the plane or space which will give us a vector field.

Fig. 3.2 Space curves defined by position vectors

To define the derivative of a vector, let O be the origin and P be the position of a moving particle at time t as shown in Fig. 3.3 The position vector of the point is given by . Now if this particle is moving with time then let Q be its position at time t+t From Fig. 3.1 it can be seen that
O

Fig. 3.3 Position and Velocity vectors

) ( The definition of the change in position is given by Since this change occurred during a small instant of time t then the motion can be expressed by the velocity

which is a vector. As t0, Q

tends to P and the chord PQ becomes the tangent at P. The general definition of the velocity vector is

Note that

is a vector in the direction of the tangent at P. It gives the velocity of

of the particle at P which is along the tangent to its path. This derivative can be found by finding the derivative of the components of the vector .

is also called the differential coefficient of with respect to t.

In a similar way we can define

as the second derivative of . Also

gives the acceleration of the particle at P. | |. ) | | ( ) | |

In this sense speed is defined as So ( )(

Theorems of Differentiation: Using the theories of vectors it can be shown that (i) (ii) ( ( ) )

(iii) (iv)

( ) ( )

where the order of the functions is not to be changed Ex: A particle moves along the curve in the direction Sol. The position vector The velocity

where

is the time. Find the component of its velocity and acceleration at time . ( ) ( )

and when t = 1, we have

Now to find the component of in the direction of orthogonal projection from chapter 1 ( | | | | )( )

we use

Now acceleration

and when t = 1, we have Component of acceleration in the direction of ( | | | | Hw1: The position vector of a particle at time is ( ) ( ) Find the condition imposed on by requiring that at time t=1, the acceleration is normal to the position vector. )( is given by )

3.2 Application: Projectile Movement


When we shoot a projectile into the air, we usually want to know beforehand how far it will go (will it reach the target?), how high it will rise, and when it will land. We get this information from the direction and magnitude of the projectiles initial velocity vector, using Newtons

second law of motion. This application is highly important to communication engineers to understand the physics and geometry of placing satellites in orbit. To derive equations for projectile motion, we assume that the projectile behaves like a particle moving in vertical coordinate plane and that the only force acting on the projectile during its flight is the constant force of gravity which always points straight down producing a downward acceleration of g. We will neglect the effects of Earths motion, air friction and gravitational force change. We will assume that the projectile was launched from the origin at time t=0 into the first quadrant with an initial velocity vo at an angle with the horizontal (| | ) (| | ) The projectiles initial position is ro as shown in Fig. 3.4 Newtons second law of motion says that the vertical downward acceleration due to gravity is g so if r is the projectiles position vector and t is the time then

This is a vector-differential equation subject to the following initial conditions (| | ) (| | )

The solution can be found by integration giving the following motion vector (| The | ) of ((| r | give ) the ) parametric
Fig. 3.4 Projectile Motion

components

equations ( ) (| | ) and ( ) (| | )

The position vector at any time t can be found and hence the position of the projectile can be simply drawn Projectile Height, Flight Time and Range To find the maximum projectile height, maximize y(t) | | (| | ) Substitute this value of t get (| | )

To find the maximum projectile range, set y=0 and solve for t to get the time of flight T ( ) (| | ) | |

Substitute this value in x(t) to get the maximum horizontal range R | | | | ( ) (| | ) Ex: A bullet projectile is fired from the origin over horizontal ground at an initial speed of 500m/s and a launch angle of 60 (a) (b) (c) Sol. (a) (( (b) , (c) (( (|
(| | | |

Where will the projectile be 10 seconds later? Find the maximum height, flight time and range of the projectile Find the speed of the projectile when it hits the ground back

(|

| )

((|

) )

( (

) )

| )

((| )

) | |

In conclusion a bullet fired in the air will reach a maximum height of less than 12km, travel a distance of less than 25km and take less than 2

minutes to fall down Hw2: Find the launch angles that will give the maximum 1. Height 2. Range 3. Flight time Hw3: A baseball is hit when it is 1 m about the ground. It leaves the bat with an initial speed of 50 m/s making an angle of 20 with the horizontal. At the instant the ball is hit, the wind was blowing horizontally opposite to the ball direction adding a component of -2.8i (m/s) to the balls initial velocity (a) (b) (c) Find the position vector of the path of the baseball How high does the baseball go when it reaches its maximum height? Assuming the ball is not caught , find its range and flight time

3.3 Arc Length, and Unit Tangent Vector


Airplanes making turns and twists are subject to very high forces due to the high accelerations. Turns that are too tight or climbs that are too steep can even breakup the airplane or spin it out of control to crash to Earth.
Airplane.wmv

Depending on the shape of the curve we can determine the sharpness of its turning. One of the features of smooth space curves is that they have a measurable length. This means that we can locate points along the curve by knowing their direct distance s along the curve from some base point or reference as shown in Fig. 3.5. In this figure the time variable describes the motion of the particle while s describes the distance traveled along this curve
Fig. 3.5 Smooth curves can be scaled like number lines

Definition: Length of a Smooth Curve The length of a smooth curve traced only once as t increases from ( ) ( ) ( ) ( ) to ( ) ( ) , is that is

Please note that the square root in this definition is the magnitude of the velocity vector = || | | so we can say || Ex: An airplane is gliding upward along the helix ? Sol. || ( ( ) ) ( ) ( ) ( ) ( to ) .

How far does the airplane travel along it path from

Fig. 3.6 Example

If we choose a base or reference point ( ), each value of t will gives us a point ( ) point. If if ( ( ) ( ) ( )) on the curve with respect to the reference , then ( ) is the positive distance from ( ) to ( ) while

, then ( ) is negative. For any value of t, the distance s(t) will be ( ( ))

called the are length parameter and will be defined as ( ) ( ( )) ( ( )) |( )|

HW4 Show that if

is a unit vector, then the arc

length parameter ( ) along the line ( ) ) from the point ( ) where

) itself

is equal to

Note: since the curve is smooth then ( ) is differentiable with respect to and the derivative is a scalar function that gives the speed of the |( )| will be an increasing function of time. That is never zero for a smooth curve Unit Tangent Vector T We already know that the velocity vector and the vector || is therefore a unit vector tangent to the curve as shown in Fig. 3.7 Definition The unit tangent vector to a smooth curve ( ) is given by ||

object

since |( )| is

is tangent to the curve

Fig. 3.7 Unit tangent vector

Ex: Find the unit tangent vector of the curve Sol. Thus
||

||

HW5: Given a counterclockwise moving phasor in 2D plane as shown in Fig. 3.8 which can be described by the position vector ( ) find the unit tangent vector to the circle ( ) ( ),

Fig. 3.8 Homework

3.4 Curvature, and Unit Normal Vector


As a particle moves along a smooth curve in the plane,

turns as the

curve bends. Since is a unit vector, its length remains constant and only its direction changes as the particle moves along the curve as shown in Fig. 3.9

Fig. 3.9 As P moves along the curve in the direction of increasing length, the unit tangent vector turns.

The rate at which T turns per unit of length along the curve is called the curvature Definition If T is the unit vector of a smooth curve, the curvature function of the curve is || The curvature is a scalar function and if | | is large, turns sharply as the particle passes through , and the curvature at is smaller Ex: On a straight line, the unit tangent vector always points in the same direction, so its components are constants. Therefore | | | | as shown in Fig. 3.10 | | | | | |

Fig. 3.10 Along a straight line, T always points in the same direction. Therefore the curvature is zero.

Ex: Find the curvature of a circle of radius a Sol. The position vector for a circle of radius a is ( ) ( ) ( ) ( ) ( ) || So Thus || ( ) ( ) ( ) ( )

( (

) )

( (

) )

Hence for any value of t || | |

There are many vectors normal to the vector but one of them is of special interest since it points in the direction in which the curve is turning. Note that will be a function of the parameter

so it can be considered

a curve of . Since has constant length (unit vector), the derivative is orthogonal to . Therefore if we divide

by its length , we obtain a

unit vector orthogonal to as shown in Fig. 3.11

Fig. 3.11 The vector dT/ds, normal to the curve, always points in the direction in which T is turning. The unit normal vector N is the direction of dT/ds

Definition At a point where 0 the principal unit normal vector for a smooth curve in the plane is Where The vector

| || |

| |

points in the direction in which turns as the curve bends.

Therefore, if we face in the direction of increasing arc length, the direction of the vector

is towards the right if turns clockwise and

toward the left if turns counterclockwise. In other words, the principal normal vector will point toward the concave side of the curve as can be seen in Fig. 3.11 Note that we can find without having to find or s Ex: Find and for the circular motion ( ) Sol. We first find , ( ) || From which we find || Notice that | | ( ) ( ) || ( ) ( ) ( ) ( )

, verifying that is orthogonal to . Notice too, that for the circular motion here, points from ( ) towards the circles center at the origin. Note: The binormal vector of a curve in space is orthogonal to both and as shown in Fig. 3.12 , a unit vector

Fig. 3.12 The mutually orthogonal vectors travelling along a curve in space

Ex: Find the curvature for the helix ) ( ) ( Then find the principal normal vector and the binormal vector Sol. The helix is shown in Fig. 3.13. First we find the velocity ( ) ( ) ( || || | | | ) [ ( ] [ ( ( ) ) ( ( )] ) ) ( )

||

Fig. 3.13 Example

| |

[ (

)]

Note that always points towards the z-axis To find the binormal vector we apply

[( ) ( ) ] HW6: For the following position vector find the unit normal, tangent and binormal vectors ( ) ( )

3.5 Gradient Vectors and Directional Derivatives


In the previous section we considered a vector function that is parameterized to become a function of a single variable . In this section we return to the concept of differentiation of functions of two or more variables but we will start with scalar functions of two variables ( Recall that the partial derivative function ( Similarly, the partial derivative ). was the rate of change of the is the rate of change if you walk in ) if you walk in some other

) if you walk in the positive x-direction with unit speed.

the positive y-direction with unit speed. We are left with the obvious question, What is the rate of change of ( direction? Let represent a vector which points in the direction of travel. The length of will be one, to reflect the fact that we are walking with unit speed. We already know two things: Rate of change of Rate of change of ( ( ) in direction of x or in the direction ) in direction of y or in the direction

It is reasonable, then, to expect that Rate of change of ( Ex: Suppose ( ) ) in direction of x or in the direction . Then what is the rate of change of (

), at

the point (2, 1), in the direction

Sol. The answer will be

The rate of change of (

) (scalar function of two or more variables),

in the direction of , is called a directional derivative, and is denoted as ( ). Hence, we have the formula

But notice that the right side of this equation can be rewritten as a dot product:

We can also write this equation shorter by coming up with a new notation for for ( . The best way is to name this last vector as ) then becomes

) and we

will call this vector the gradient of f, and denote it as

. The equation

( ) is a vector as long as

From this definition we can now see that ( ) is a scalar function and that ( Sol. ( ) ) )

Ex: Find the gradient of the function (

at the point (x, y)

What Do the Directional Derivative and the Gradient Mean? The gradient at the point (1, 1) in the previous example, then, would be ( ) Remember that the rate of change of a function is also the slope of a tangent line to its graph, as long as you are traveling with speed one. Heres a nice way to think about the situation. Suppose you are climbing a mountain, and you have a good trail map in your hands. Let ( ( ) be your coordinates when you locate yourself on the map. The function ) is your elevation at that point. Now turn your body to face the direction (on your map). If you sight up or down so that your gaze just grazes the mountainside then you are looking along the tangent line

whose slope is given by

) as shown in Fig. .

Fig. 3.14 Interpretation of the directional derivative

Recall that the dot product is given by the product of the magnitudes of the two vectors, times the cosine of the angle between them. If we fix the point we are at then way to change

is a fixed vector. If ||

, then the only . is

is to change the angle between and

The largest this quantity can be is when the value of cos is largest. This happens when = 0. We conclude that the largest value of given by

|||

and that this value is attained when (the direction we are facing) coincides with the direction of . In other words, if you are on the mountainside and you want to face directly uphill you should point yourself in the direction of the gradient vector. When you do this and sight along the mountainside the slope you see is the magnitude of the gradient vector. (Q) What happens if you are in a valley and you want to face the bottom of the valley? EX: Let ( ) find the largest directional derivative of this function at the point (2,3) and the largest slope for any tangent line at

this point. Sol. At the point (x, y) the gradient vector is ( ) ( ) ) ( ) . The largest slope So, at the point (2, 3) we have | ( )| (

of any tangent line to the graph at the point (2, 3) is then given by

(Q) What if you were standing on the mountainside and wanted to face the direction you would have to travel to keep your elevation constant? In other words, how would you find the direction of your level curve? (A) If you were facing such a direction you would be looking along a horizontal line, i.e., a line whose slope is zero. The only way for to be zero is for and ) to be perpendicular. . In the previous example we saw ( )

EX: Suppose again ( would be

. A vector which points in a direction perpendicular to this (Check this!). Hence, this vector is tangent to a level curve at the point (2, 3). HW7: For each of the functions questions: Find the gradient vector . Find a unit vector that points in the direction of the maximum rate of change at the point (1, 1). Find the largest slope of any tangent line at (1, 1). Find a unit vector that lies in a line tangent to a level curve through (1, 1). 1. 2. 3. ( ). ) at the point (1, 1) in the direction Calculate the rate of change of ( ( ) below answer the following

3.6 Maxima, Minima, and Saddles


At a local maximum or a local minimum of a graph the tangent plane is horizontal. Another way to say this is that the slope of every tangent line to a local maximum or minimum is zero. But recall that the slope of a tangent line at ( ) in direction is given by . The only way for this to be zero for every possible is if EX: Show that ( Sol. The gradient is If x and y are nonzero then this is not the zero vector. This tells us that there is a direction where the slopes of tangent lines are nonzero, and hence we cannot be at a local maximum or minimum. Unfortunately, just because the gradient is the zero vector it does not necessarily mean that there is a local maximum or minimum. EX: The gradient of Sol. The function ( ( ) and ( ) is at ) is the zero vector

cannot have a maximum or minimum at any

point where x and y are nonzero.

the origin. What is the difference between the two? ) has a minimum there, while ( ) has a saddle. In the first-semester we learned to detect local maxima and minima by a second-derivative test. We would like to do the same thing here. The problem, of course, is that there are four second partial derivatives! To keep track of all this information we often write them in a matrix, as follows:

[ ] Now we examine this matrix for several functions whose graph is familiar. Each of these has a gradient vector equal to zero at the origin. 1. . This function has a local minimum at the origin. The matrix of second partials is

* 2. partials is * Let us take other examples 3.

. This function has a saddle at the origin. The matrix of second +

. This function has a maximum at the origin. The matrix of * +

second partials is 4. . This function also has a saddle at the origin. The matrix of second * +

partials is The first and third examples give us a clue as to the quantity we would like to look at. Consider the product of the upper-left and lower-right entries of the matrix. For the maximum and minimum above this quantity is positive and for the first of the above saddles it is negative. However, this alone would not be enough to distinguish maxima and minima from saddles, as the second of the saddles shows. To compensate we must subtract the product of the upper-right and lower-left entries, yielding the formula

However, since the mixed partials are equal we can shorten this to ( )

This is indeed the right quantity to look at, in the sense that if it is greater than zero you have a maximum or minimum, and if it is less than zero you have a saddle. Unfortunately, if it is zero you have no information; you may be at a maximum, minimum, saddle, or something much more bizarre. Nonetheless, we will single this out as our first test.

EX: Find the location of all saddles of the function

Sol. First, we will need to narrow down the possibilities by finding the critical points. To do this we find the gradient. ( ) ( ) Setting this equal to the vector

gives us the system of equations

The first equation tells us that equation then gives Either (and hence

. Plugging this into the second

) or we can divide both sides of this

equation by y to get Solving then gives us Hence, we have critical points at (0, 0) and (-2/9, 2/9). To determine which of these are saddles we compute the matrix of second partials: [ [ And so ( ) ( ) ) We now check each critical point: ( ) ( ) ( ] ]

When

it would be nice to have a second test to determine ( ) and ( )

whether ( .

) is a local maximum or a local minimum. Such a test can be

easily guessed from our typical examples, Notice that in both cases and

have the same sign. When this sign

is positive we have a local minimum and when it is negative we have a local maximum. This is precisely our second test.

EX: Recall the function

from the previous

example. We found critical points at (0, 0) and (-2/9,2/9), and determined that at (-2/9,2/9) there was a local maximum or a local minimum. To determine which we need only look at as shown in Fig. 3.15. . Since this was 2, and 2 0, we conclude that at this critical point there is a local minimum

-2

-4 1

0.5

-0.5

-1

-1

-0.5

0.5

Fig. 3.15 Example

It is important to keep in mind that if information about the nature of ( example. EX: Consider the following functions: 1. ( 2. ( 3. ( ) ) ) .

then we have no

). We illustrate this in the next

In each case the only critical point is at (0, 0) and maximum, and in the third there is a saddle.

. But at (0,

0) in the first case there is a local minimum, in the second there is a local

HW8: Find the local maxima, minima, and saddles of the following functions. 1. 2.

HW9: For the function sin(x + y) show that D(x, y) = 0 at every point (x, y). Does this function have maxima, minima, or saddles? HW10 If, for some point ( then show that instead of ( ( ) ) you know ( ) and ( ( ) )

. (This tells us that you may use

) when distinguishing maxima from minima.)

3.7 Vector Fields and Gradients


Definition A vector field is simply a choice of vector for each point. So, for example, a vector field on R2 would have some vector at the point (1, 2), some other vector at the point (1, 1), etc. We often draw vector fields by picking a few points and drawing the vector based at those points as shown in Fig. 3.16.

Fig. 3.16 Vector Field

More formally, a vector field is a function from R2 to R2. What goes in to this function are the coordinates of the point where you are. What

comes out are the components of the vector at that point. EX: Consider the vector field field contains the vector the vector . At the point (1, 1) this vector . At the point (2, 4) it contains . If we use a computer to draw it we would

see something like what is shown in Fig. 3.17.

Fig. 3.17 Example

EX: The vector field

is shown in Fig. 3.18.

Fig. 3.18 Example

HW11 Sketch the following vector fields. 1. 2. 3. We have already seen many vector fields, although we did not use this language. Whenever we take a function f and compute its gradient f at a point we get a vector. The set of all such vectors is then a vector field, which we now call grad f. EX: Suppose ( is ( ) ) . Then the gradient of ( ) at the point ( ) . If we draw this vector at various values of x and

y, we get the picture depicted below in Fig. 3.19.

Fig. 3.19 Example

EX: Let (

. Then

HW12 For each of the following functions, f, compute f. 1. ( 2. ( 3. ( ) ) )

3.8 Divergence
In the previous section, we saw that the gradient operator gives us a way to take a function ( ) and get a vector field. In this section, we explore a way to take a vector field and get a function. Eventually, we will see that the value of this function at a point is a measure of how much the vector field is spreading out there. Definition Suppose ( ) ( ( ) ( ) ) ( ( ) is a vector field V on R3 (that is ) ). Then we define the divergence of

V, Div V, to be the function

Note that the first term is associated with the first component of V, the second term with the second component, and the third term with the third component. This, and the fact that the terms are being added, should remind you of the dot product. This gives us a purely notational way to remember how to calculate the divergence of a vector field. We let denote the vector

This, of course, is only a vector in a notational sense. But if we suspend our disbelief for a moment and allow such absurdities, we can write the formula for the divergence of a vector field in a very compact way: The is treated as a vector operator and is called del operator or nabla. Generally it is not a real vector but describes a vector operation. Therefore it has the same properties of vectors. EX: Let V be the vector field ( )

. Then the divergence of V is calculated as follows:

What does the Divergence of a Vector Function mean? Divergence is a vector operator that measures the magnitude of a vector field's source or sink at a given point per unit volume, in terms of a signed scalar. It simply measures the rate at which "density" exits a given region of space. In the absence of creation or destruction of matter, the density within a region of space can change only by having it flow into a sink or out of a source in the region. Practically, the divergence measures the volume density of the outward flux of a vector field from a very small volume around a given point. For example, consider air as it is heated or cooled. The relevant vector field for this example is the velocity of the moving air at a point. If air is heated in a region it will expand in all directions such that the velocity field points outward from that region. Therefore the divergence of the velocity field in that region would have a positive value, as the region is a source. If the air cools and contracts, the divergence is negative and the region is called a sink. Figure 3.20a shows two regions in a vector field one representing a source with vectors emerging outward and one represents a sink with vectors converging inward. A contour plot of the divergence shows that the source region has a positive divergence while the sink region has a negative divergence. Fig. 3.20b shows the divergence function in 3D. If the divergence is nonzero at some point then there must be a source or sink at that position. A vector field with constant zero divergence is called incompressible or solenoidal in this case, no net flow can occur across any closed surface.

0.5

-0.5

-1 -2 -1.5 -1 -0.5
(a)

0.5

1.5

3 2 1 0 -1 -2 -3 1 0.5 0 -0.5 -1 -2 -1 0 1 2

(b) Fig. 3.20 A vector field and its divergence

Note that by measuring the net flux of content passing through a surface surrounding the region of space; it is immediately possible to say

how the density of the interior has changed. The intuition that the sum of all sources minus the sum of all sinks should give the net flow outwards of a region is the principle of divergence theorem as will be shown later. Applications of Divergence (Q) Give some examples of where will you see divergence. As discussed before divergence is the outflow of flux from a very small closed surface area (per unit volume). It will be positive if the fluid is expanding or if it is being supplied by a source external to the field. It will be negative if the fluid is shrinking or if it being absorbed by a sink. If there is no loss or gain of the fluid anywhere then the divergence will be zero { The applications of divergence in physics are very wide 1. Electromagnetic fields: Gausss law electric The sources or sinks of the electric field in any material or medium are the electric charges. If the volume was very small the divergence would be positive for positive charge and negative for negative charge

where the volume charge density is the total charge divided by the volume of the medium and the permittivity is a constant describing how the material or medium condenses or spreads the field lines 2. Electromagnetic fields: Gausss law magnetic The sources and sinks of the magnetic field in any material or medium cancel each other. No matter how small the volume was, it will always contain both the north and south poles. Thus the divergence cannot have any value other than zero because the positive sources are canceled by the negative sinks

3. Fluid mechanics and electromagnetic fields: Continuity equation Consider velocity of the air in a tire that has just been punctured by a nail, the air is expanding as the pressure drops, and consequently there is a net outflow from any closed surface lying within the tire. The divergence of this velocity is therefore greater than zero. The expanding air produces a positive divergence of the velocity, and each interior point may be considered a source. The continuity equation of fluid mechanics states that the rate at which density decreases in a very small volume element of fluid is ( where ) proportional to the flux of fluid flowing away from the element,

is the vector field of fluid velocity. In electromagnetic

fields the continuity equation is written as

where

is the current density vector field


( )

Ex: Assume the air velocity around a freezer is given by 1. Find the gradient particles 2. Find the divergence of the vector field proportional to the temperature of the air Sol. 1. 2.
( )

(slopes) vector field of the velocity of air (which will be )


( ) ) ( ( ) )

(
(

Note that this last operation denoted by

is called the Laplacian and is

HW13 Compute the divergence of the following vector fields. 1.

2. 3. HW14 Given that the voltage function in free space is given by , find the associated electric field ( distribution of the charges per unit volume ( HW15 Let ( ) be a function. What is ? ) ) then find the

3.9 Curl
There are many useful ways to apply partial derivatives to vector fields. We have already seen that gradients give us a way to take a function and get a vector field representing slopes of this function in the direction of the peak. Then we saw that divergence is a way to take a vector field and get a function representing the sources and sinks of this vector field. Here, we define a way to use partial derivatives to transform one vector field into another. First, recall our notational absurdity, . In the previous section, we defined a new operation by using ( in a dot product. Here, we define an operation called curl by using ( ) ( ) ( ) ) ( | ) ( | ) . Then we define ( ) ( ) ( ) in a cross product. As before, let

) (

) (

) ( )

EX: Let V be the vector field

. Then the curl of V is calculated as follows: | | | | ( ) ( )

Later we will see that the curl of a vector field measures how much it twists at each point.

What does the Curl of a Vector Function mean? The curl of V at a point in a fluid is a measure of the rotation of the fluid around this point. If there is no rotation of fluid anywhere then . Such a vector field is said to be irrotational or conservative. For a 2D flow with V representing the fluid velocity, is perpendicular to the motion and represents the direction of axis of rotation. It measures the maximum circulation or angular momentum of the vector field at each point per unit area when this area becomes very small. The direction of the curl is the axis of rotation, as determined by the right-hand rule, and the magnitude of the curl is the magnitude of rotation Suppose the vector field describes the velocity field of a fluid flow (such as a large tank of liquid or gas) and a paddle wheel is located within the fluid or gas (the centre of the wheel being fixed at a certain point). The fluid flowing past the wheel will make it rotate. The rotation axis (oriented according to the right hand rule) points in the direction of the curl of the field at the centre of the wheel, and the angular speed of the rotation is half the magnitude of the curl at this point. To find out if the curl has a value for some vector field we can imagine the above paddle wheel placed in the vector field and see if it rotates or not. If it does then the curl has value. Otherwise the curl will be zero. Ex: Consider the vector field shown in Fig. 3.21 ( ) . This vector field is

Fig. 3.21 Example

Simply by visual inspection, we can see that the field is rotating. If we place a paddle wheel anywhere, we see immediately its tendency to rotate clockwise. Using the right-hand rule, we expect the curl to be into the page. If we are to keep a right-handed coordinate system, into the page will be in the negative z direction. If we calculate the curl: | |

which is indeed in the negative z direction, as expected. In this case, the curl is actually a constant, irrespective of position. The "amount" of rotation in the above vector field is the same at any point (x, y). Ex: Consider the vector field shown in Fig. 3.22 ( ) . This vector field is

Fig. 3.22 Example

We might not see any rotation initially, but if we closely look at the right, we see a larger field at, say, x=4 than at x=3. Intuitively, if we placed a small paddle wheel there, the larger "current" on its right side would cause the paddlewheel to rotate clockwise, which corresponds to a curl in the negative z direction. By contrast, if we look at a point on the left and placed a small paddle wheel there, the larger "current" on its left side would cause the paddlewheel to rotate counterclockwise, which corresponds to a curl in the positive z direction. Let's check out our guess by doing the math: | |

Indeed the curl is in the positive z direction for negative x and in the negative z direction for positive x, as expected. The plot of the curl in this case is shown in Fig. 3.23

Fig. 3.23 Plot of the curl in 3D

Applications of Curl (Q) Give an example of where will you see curl. Electromagnetic fields: Faradays and Amperes laws for time varying fields Rotation of electric field in space causes a perpendicular time varying magnetic filed

Rotation of the magnetic field in space causes a perpendicular time varying electric filed plus some current in any nearby conductor

Properties of Divergence, Gradient and Curl ( ) 1. ( ) ( ) ( ) 2. ( ) ( ) ( ) 3. ( ) ( ) ( ) 4. ( ) ( ) ( 5. ( ) ( ) 6. Important Identities to Remember ( ( Ex: Verify that Sol. ( ) ) )

for the function (

| | ( ) (

| | ) ( )

Helmholtz Decomposition and Reconstruction Any vector field V on a bounded domain C in R3, which is twice continuously differentiable can be decomposed into a curl-free component and a divergence-free component. That is V can be written as for some function ( and the vector field for some operator ( ) So if the curl and divergence of a vector are known then the vector is known HW16 Show that ( ) ( ) ( ) ) and some vector field . and the vector field is zero at infinity then can be reconstructed by ( ( )) ( ( ))

Similarly, if we know the function

HW17: Find the curl of each of the following vector fields. 1. 2. HW18: Verify that . ( ) for the vector field

3.10 Integrating Vector Fields


In the previous sections we introduced the concept of a vector field, and looked at various concepts of differentiation in relation to these objects. In this section, we look at ways to integrate a vector field. From calculus you should suspect that when we combine integration and differentiation something amazing will happen, akin to the Fundamental Theorem of Calculus. This is the topic of the subsequent sections.

3.10.1. Line Integrals


We will always integrate a vector field over a domain that can be parameterized by a line, surface, or volume. In this section, we look at the first of these situations, integrals over parameterized curves. These will be called line integrals. However, the definition of the integral of a vector field over a parameterized curve is a bit different than the definition of a function over the same curve. Before giving the definition, it may help to give a physical motivation for line integrals. If you drop a leaf into a flowing stream and follow it. Eventually, you may see it encounter a stick caught on some rocks. If the stick is perpendicular to the flow of water, the leaf will get stuck. If the stick is parallel to the flow, then the leaf rushes by as the force of the flowing water pushes it along. But if the stick is at some angle to the flow the leaf may hit it, and then slowly work its way past.

Fig. 3.24 Physical Motivation for Line Integral

One can explain this by looking at the force the water can exert on the leaf. The leaf can only travel in a direction parallel to the stick, which we can represent as a vector, . The water flowing under the stick exerts . If some force on the leaf, which can also be represented as a vector, force

these vectors are perpendicular then the leaf does not move, so the net on the leaf must be zero. On the other hand, the leaf moves the and are parallel, so must be greatest in this is proportional to the dot fastest when product, .

situation. It should be no surprise, then, that

Fig. 3.25 Interpretation of the Force

Note that in the above figure 3.25, there were two choices for the vector . One of these is shown in the figure, and the other points in the

exact opposite direction (but still along the stick). If the other choice was made, then the value of would have exactly the opposite sign. So which is the correct choice mathematically? There is no right answer to this. We just have to declare, before we do any problem, which is correct depending on the physical situation. Such a declaration is called an orientation. We will say more about this idea later. For now, just note that if the wrong choice was made then we can correct things by changing the sign of our final answer. Now we suppose that somehow the leaf is only moving on some parameterized curve, C, as the water rushes past. Sometimes the water may be perpendicular to the curve, and sometimes it may be parallel. The problem is to evaluate the total force that the water exerts on the leaf as it travels along the curve. (If the leaf always moves with unit speed then this is equal to the work done by the water.) Let ( ) be a parameterization of C. Let be a vector field that gives the direction and speed of the water at each point of the stream. The direction that the leaf is moving at the point ( ) is the tangential vector . Let ( ( )) be the vector of the vector field that is based at the point ( ). Then the quantity we want to evaluate is as Our answer, unfortunately, is still incomplete. Recall that at some point we must make a choice, called an orientation. This amounts to deciding if the vector points the right way or the wrong way. An orientation can be denoted pictorially just by an arrow along C, showing the correct choice. But this does not mean that to find the integral, we must choose a parameterization whose derivative points the right way. If it does not then we just need to change the sign of our final answer. ( ) ( ) ( ( ))

This integral is called the line integral of

over C, and is often written

EX: Let

. Let C be the oriented curve shown in Fig. 3.26.

Fig. 3.26 Example

The orientation is given by the direction of the arrow in the figure. The curve is parameterized by ( ) one can easily check that
( )

. At the point ( ) . This vector points in the

direction opposite to the orientation. This just means that after we integrate we will have to change the sign of our final answer. We now calculate the line integral of that ( ( )) and ( )( ) ( ) along C using ( ). First, note

* +

As the orientation disagrees with obtaining our final answer, 64.4.

( ) we must now change the sign,

An easy way to give an orientation on a curve parameterized by a function ( ) is to simply declare that points in the right direction. In this case, without any we may go ahead and use ( ) to evaluate the integral of

further considerations. In this case, we say the orientation of C is the

one induced by ( ). EX: Let C be the curve parameterized by ( ) with the orientation induced by ( ). Let C. Sol. First, note that Next, observe that . Integrate over

We now integrate [ | |]

As the orientation of C is chosen to agree with ( ) we do not need to worry about changing the sign of our answer. The connection between line integrals of vector fields and line integrals of functions Let ( ) denote a function and C a curve parameterized by ( ). Let ( ) of C, the vector and has be a vector field chosen so that at the point magnitude of ( ( )). Then we have ( ( )) The definition of the line integral of ( ( )) ( ( ))

( ( )) is tangent to C (i.e., it points in the same direction as

over C then gives ( ( )) | | |

| |

| |

( ( )) | |

( ( )) |

This last equality is precisely the definition of the line integral of a scalar function over C.

Ex: Find the line integral of the function ( ) Sol.

over the curve

| ( ( )) | |

( )( ) * +

HW19* Calculate the line integral of oriented counterclockwise. Hint: ( ) HW20 Let ( ) ( ) Calculate the line integral of HW21* Let orientation. 1. Calculate the integral of the vector field 2. Calculate ( ( )) ( ( )). ( ) by

over the circle of radius 1,

. Let C be the curve parameterized by with the induced orientation. over C.

. Let C be the portion of the parabola ( ) with over C. the induced

parameterized

HW22* Let C be the subset of the graph of (oriented away from the origin). Let Integrate over C.

where .

be the vector field

3.10.2. Surface Integrals


In the previous section we defined line integrals of vector fields and line integrals of functions. By analogy we will now define surface integrals of vector fields and surface integrals of functions. The physical motivation for surface integrals of vector fields can again be seen by looking at a river of flowing water. Imagine a net placed vertically in the river as shown in Fig. 3.27. Suppose you wanted to calculate the amount of water flowing through the net at some particular

moment in time.

Fig. 3.27 Physical Motivation for Surface Integral

Let

denote a vector representing the flow of water at some point of be a vector which is perpendicular to the net. If the net is is perpendicular to . The most water passes through the and are

the net. Let situation,

parallel to the direction of flow, then no water passes through it. In this net when the net is perpendicular to the flow. In this case, through the net is proportional to as shown in Fig. 3.28.

parallel. It stands to reason, then, that the amount of water flowing

Fig. 3.28 Interpretation of net flow

Once again there is a technical problem. Why did we draw the vector the figure the way we did? If the only condition on perpendicular to the net, then will discuss this more shortly. Now let S be a surface (representing the net), parameterized by (

in

is that it is

could point in the exact opposite

direction. The correct choice will again depend on the orientation. We ).

The vectors water flow, and

and ( (

are both tangent to S. Hence, the vector )) the vector of ( ( )) ( at the point ( ( ( )) )

is

perpendicular, or normal, to S. Let

be a vector field representing the ). Then the

total amount of water flowing through S is given by the integral

We call this the surface integral of

over S, and denote it by

But this formula isnt the whole story. We still must deal with the issue of orientations. Otherwise, people evaluating may get different answers, depending on which parameterization of S they use to evaluate the integral. One way to give an orientation is to say which way is up at each point of S. This can be done by giving a vector O which is perpendicular to S at some point. Our parameterization the choice of orientation if the vector agrees with points in the same direction

as O. If our parameterization does not agree with the specified orientation then we can change the sign our final answer. EX: Let by ( ) ( . Let S be the portion of a cylinder parameterized ) . The vector

defines an orientation on S at the point (1, 0, 0). We now compute the integral of the partials, over the surface S with this orientation. First, we compute

And so, | Now from ( ) ( | ) we notice that

. So (

( (

))

.Now notice that the point (1, 0, 0) = is equal to at this point. This is over S we will have

). The vector

exactly opposite to the specified orientation of S. Hence, if we use the parameterization ( We now integrate ( ( )) ( ) ) to compute the integral of to remember to change the sign of our final answer.

As the parameterization ( the correct answer is HW23 Let . . Integrate over the unit sphere, with at the point (1, 0, 0). Hint: a unit ) . Integrate over the surface S ) disagreed with the specified orientation

orientation given by the vector function ( HW24 Let ) (

spherical surface in spherical coordinates is parameterizes by the

parameterized by ( ) Use the induced orientation. HW25 Let S be the surface given by the following parameterization: ( ) ( ) (Note that this is not quite spherical coordinates.) Integrate over S the

vector field

HW26* Let vector

Let S denote the intersection of the unit at the point (0, 1, 0). Compute ( )

sphere with the positive octant. An orientation on S is given by the

3.11 Integration Theorems


We noticed before from HW21 that for a curve C parameterized by r(t) over the range ( ( )) ( ( ))

This was no coincidence and in fact it holds true for any scalar function.

3.11.1 Path Independence Theorem


Suppose r(t) is a parameterization of C with domain [a, b] which agrees with the orientation on C. Then for any function ( ( )) we have

( ( ))

Often this is referred to as the path independence of line integrals of gradient fields. This is because it says that the result of a line integral of a gradient field only depends on the endpoints of the curve, and not the path used to get from one endpoint to the other. EX: Suppose ( ) . Let C be the top half of the unit circle,

oriented counterclockwise. We compute All we really have to know is the endpoints of the curve C. The first is (1, 0) and the second is (1, 0). (Which one is which is determined by the orientation on C.) Hence, ( ) ( )

Applications of Path Independence of Line Integrals of Gradient Fields There is an important application of the independence of path of line integrals of gradient fields that may be familiar. Recall that a line integral of a vector field says something about how much work you have to do to move an object along a curve C in the presence of a force W. Lets say you want to know how much work you have to do against the force of gravity to get a heavy package up a mountain. Suppose the mountain is represented by the graph of a function f(x, y). That is, at the map coordinates (x, y), the function f(x, y) gives you your altitude. Then the force of gravity at the point (x, y) is proportional to the vector f(x, y) (i.e., the steeper the mountain, the harder you have to work to get up it). The constant of proportionality is the objects weight = mass9.81. Lets suppose you have identified your route on a map. You start at the point (x0, y0), follow some curve C, and end up at (x1, y1). Then the work you have to do to overcome gravity is proportional to ( ) ( )

Notice that the result is just the difference in elevation between your beginning and ending point, and doesnt matter what path you take to get from one to the other as shown in Fig. 3.29!

Fig. 3.29 Path independence of the line integral of the gradient

Another example is the potential difference between two charge objects

( ( )

( ))

EX: Let the potential function near a charged surface be given by the function charge 2. Find the voltage needed to move a unit charge along the curve C parameterized by orientation Sol. 1. ( ) 2. We would like to integrate E over the curve C. This integral can be done directly, but the wise reader will notice that ( ) with the induced . 1. Find the associated electric field that would be applied on a point

Hence, we may use the independence of path of line integrals: Or ( ( )) ( ( )) ( ) | ( ) | ( ( )) ( ( ))

HW27 Let

. Suppose C is some curve that goes from (1, 0, 1)

to (1, 1, 1). Calculate the integral of W over C. Hint: Assume W is the gradient of another function e.g. f=x+y+z HW28 Let C be the curve pictured beside. Let function. . Calculate . Hint: Assume W is the gradient of another

HW29 Suppose (

. Let C be a curve that is

starts at (2,1) and ends at some other point. Show that (strictly) larger than zero.

HW30 Suppose C is a closed curve, i.e., one whose beginning and ending points are the same. Show that the integral of any gradient field over C is zero. HW31 Derive the Fundamental Theorem of Calculus from the

independence of path of line integrals of gradient fields. (Hint: Begin by letting C be a curve in R1 from a to b.) HW32 In this section we saw that the work you must do to carry a package halfway up a mountain is just proportional to the difference in your starting and ending altitudes, and does not depend on the path you take. But it certainly seems like if you started at the bottom of a mountain, went to the top, and then came down to the halfway point, youd be doing more work than if you just went straight up to the halfway point. Explain.

3.11.2 Greens and Stokes Theorems


The next theorem that relates derivatives, integrals, and vector fields is special to R2. That is this theorem is special only to vector fields of the form ( ( ) ( ) ( ) . Recall that the curl of a vector field ) in R3 is defined to be the quantity | | ) (

Although the curl is only defined for vector fields in R3, we can use it to define a natural (and important!) operation on vector fields in R2. First, given a vector field vector field on R3: ( ( ) ( ) ( ) ) on R2 we can easily create a . We will continue to call this new

vector field W. We now compute its curl: | ( Hence, | | | is a function on ) ( ) | ( ( ) ( ) )

So, if W is a vector field on R2 then the result of |

R2. This function acts much like the derivative of W. We now ask, what happens when we integrate the result of such a derivative? That is, if Q is some domain (some area defined by a closed curve) in R2, then what can we say about | | ( )

For simplicity, lets assume here that the domain of integration Q is a rectangle defined in the xy-plane by x=x0 (left side), x=xn (right side), y=y0 (bottom) and y=ym (top) as shown in Fig. 3.30 with the indicated orientations.

Fig. 3.30 The Rectangle Q

If we carry out the above integration around the rectangle we will see that | | ( ) ( ( ) ( ( ) ) )

)| ( (

) )

)| ( ) ( )

The first integral is actually the line integral of parameterized by the curve ( ) ( ) over the curve . First ( ) ( )

on the right side in the shown ( ( ) ( ).

orientation. To prove this lets try to find the line integral of Next, we find integral. ( ) and then find ). Finally we

integrate the result over the range of y which will give us the first

The other three terms in the sum above give similar integrals over the remaining sides of Q. You can also notice that the minus integrals will have opposite orientations to the ones shown in Fig. 3.30. Therefore | | ( )

The sides R, L, T , and B of Q, with the orientations as indicated in Fig. 3.30, when taken together are referred to as its boundary. The usual notation for this is Q. Using this notation we can write our conclusion much more compactly: | | ( )

where the symbol

refers to a closed curve around the region Q.

This final equation is known as Greens Theorem. EX: Let Q be the rectangle in the plane with corners at (0, 0), (2, 0), (0, 3), and (2, 3). Let . We will use Greens Theorem to

evaluate the integral of W over Q. ( | ) | ( )

Significance of Greens Theorem Greens Theorem enables us to see the geometric significance of the value of at a point (x0, y0). Let Q denote a very small rectangle is roughly around this point. Then at each point of Q the value of constant, and equal to its value at (x0, y0). Hence, ( ) ( ( ( ) ) ( ) ( ) )

( ( But Greens Theorem says (

( (

) )

) ) ( )

Putting these together gives us ( So, the function ( ) ( ) ) ( )

is a measure of the field around any small region

per unit area. In other words it is a measure of the circulation of W around each point. Practically one can experience such a function in real life. Many people enjoy going tubingfloating down a river in an inner tube. Suppose you are tubing and decided you want to stop somewhere to enjoy the scenery, so you drop an anchor. Then you notice that the water on your left is rushing past you faster than the water on your right. What happens? Your inner tube starts to turn. This turning is a measure

of the strength of

as shown in Fig. 3.31.

Fig. 3.31 Interpretation of Greens Theorem

HW33 Let Q be the rectangle {(x, y)|0 x 1, 0 y 1}. Use Greens Theorem to evaluate the integral of over Q.

HW34 Let C be the curve pictured below. Show that the integral of over C does not depend on b.

Now we have seen application of Greens theorem on a rectangle. But what happen if we apply it on a region with inner boundary and outer boundary. We begin by examining what happens when we look at Greens Theorem applied to two neighboring rectangles as shown in Fig. 3.32.

Fig. 3.32 Application of Greens Theorem to neighboring rectangles

The integral of W along R1 will cancel with the integral of W along L2. In general, we may use the following rule-of-thumb to figure out the orientation on each loop of its boundary: If Q is a connected region then the outermost loop of its boundary is oriented counterclockwise and all other loops of the boundary are oriented clockwise as shown in Fig. 3.33.

Fig. 3.33 Application of Greens Theorem to a region of multiple loops

EX: Use Greens Theorem to integrate the function x2 + y2 over the inside of the unit circle. Sol. If we denote this region as Q, then the boundary of Q is the unit circle itself, with a counterclockwise orientation. This can be parameterized in the usual way by ( ) Greens Theorem we must find functions f and g so that ( ) .To use

A suitable choice for g(x, y) can be found by integrating y2 with respect

to x, yielding the function xy2. Similarly, we may find f(x, y) by integrating x2 with respect to y, yielding x2y. So . Finally, we integrate ( ) ( ) ( )

( ( ))

Verification Using Mathcad

To verify the above example using Mathcad we can easily find the integral on the right hand side On the left hand side we need to specify the limits of the double integral. Since the integral is over the unity circle which can be written as So Therefore the double integral can be written as

HW35: Let following: ( ) (

. Let )

be the region parameterized by the . Calculate

HW36: Calculate the area enclosed by the unit circle by integrating some vector field around its boundary. HW37: 1. Suppose ( ) ( ) is a vector field which is defined then show that the everywhere except at (0, 0). If counterclockwise, is the same. 2. If the same. 3. Calculate the integral of HW38: ( ) . 1. Use Greens Theorem to show that 2. Let C be the horizontal segment connecting (1, 0) to (1, 0). Calculate 3. Use your previous answers to determine the integral of W over the top half of the unit circle (oriented counterclockwise). Stokes Theorem In the previous discussions, we saw that if W is a vector field in R2, then we can view |W| as a kind of derivative. When we integrated this derivative we saw something special happen, namely Greens Theorem: | | ( Let ) be the region over the unit circle. parameterized . Suppose by then show that the integral of W along every circle centered on the origin, oriented counterclockwise, is

integral of W along every circle centered on the origin, oriented

We now move our attention to R3, and explore a similar phenomenon. Suppose now W is a vector field in R3, and S is a surface. Then we wish to explore A reasonable guess, based on our experience from the Greens theorem, would be ( ( )) ( )

This turns out to be the case, and is called Stokes Theorem. We will not prove it here, as the proof is extremely similar to that of Greens Theorem. One potential complication in using Stokes Theorem is determining the boundary of the surface S in question. To get the proper orientation on S you need to know the orientation of S. Recall that this is often given by an outward pointing normal vector, n. To get the orientation on the boundary, we use the right-hand rule. To do this point the thumb of your right hand in the direction of n. Your fingers will then curl in the sense that determines the orientation on the boundary as shown in Fig. 3. 34.

Fig. 3.34 Finding the orientation of a surface

EX: Let S denote the top half of the unit sphere, with orientation given by the normal vector 1, 0, 0 at the point (1, 0, 0). We use Stokes Theorem to integrate the curl of the vector field y, x, 0 over S.

First, note that Stokes Theorem says that the answer will be the same as the integral of y, x, 0 around S. The boundary of S (with proper orientation) is parameterized by ( ) we may integrate ( ( Verification Using Mathcad To verify the above example using Mathcad we can easily find the integral on the right hand side ( ) ) ) . Thus,

On the left hand side we need to specify the limits of the double integral. Since the integral is over the top unit sphere which can be written in spherical coordinates as The element of area in spherical coordinates can be written as Where is the unit vector in the direction of Next we need to express easy since yields normal to the surface. in spherical coordinates which is . Converting this to spherical coordinated

Therefore the double integral can be written after finding the dot product to be EX: Let S denote the portion of the paraboloid z = 2 x2 y2 that lies above the plane z = 1, with an orientation determined by an upward pointing normal. Let . We will use Stokes Theorem

indirectly to find First, let D be the disk in the plane z = 1 bounded by the unit circle, with orientation given by an upward pointing normal. Then S = D. Stokes Theorem says that the integral of W over both D and S is equal to the line integral of W over S. So, to get an answer to the original problem we may evaluate the integral ofW over D instead of S. To do the integral, note that D is given by check that ( ) ( ) . So, on the plane z = 1 we have P = W = cos 1, sin 1, 0. A parameterization for

. You may

We now integrate ( ( )) ( )

HW39: Let ( ) ( ) orientation. Calculate

. Let S be the surface parameterized by with the induced

HW40: Suppose ( ) ( ) and S is a region in 3D that lies

in the xy-plane. Show that Stokes Theorem applied to W and S is equivalent to Greens Theorem. HW41: Let S be the portion of the cylinder x2 + y2 = 1 that lies between the planes z = 0 and z = 1, with orientation given by the normal vector at the point (1, 0, 0). Let integral of over S. . Calculate the

HW 42: If W is a vector field defined on all of R3, then show that the integral ofW over the unit sphere is zero.

3.11.3 Gauss Theorem


We know that one way to differentiate a vector field is to take its curl. It is then not surprising that the integral of the curl of a vector field should be special. Another way to differentiate a vector field in 3 dimensions or R3 is to take its divergence. In this section, we explore what happens when we integrate the divergence of a vector field. Suppose ( ) ( ) ( ) Recall the definition of divergence:

The result is a function on R3. We may thus integrate this function over volumes V: As in the previous sections, we might guess that there is a relationship between this and the integral of W over the boundary of V:

This equality is in fact true, and is known as Gauss Divergence Theorem. The proof is again similar to the proof of Greens Theorem. We divide the region into three dimensional small (infinitesimal) cubes in V, and we find that the integral of over each cube is approximately equal to the integral of W over the boundary of each cube (faces of the cube), with suitable orientations. But faces of cubes inherit opposite orientations from neighboring cubes, so in the sum all that is left are the faces of the cubes on the boundary of V. To properly orient the boundary of V (V), we simply choose a normal vector that points out of V. EX: Let . We would like to find the value of the integral of

over the volume V bounded by the unit sphere. According to Gauss Theorem, this is equal to the integral of W over the unit sphere S. To evaluate this, we first parameterize the unit sphere in the usual way with spherical coordinates: ( ) ( Now we compute )

To check orientations, note that at to

) this vector is equal

which does indeed point out of V. Hence, we do not need to

worry about changing the sign of an integral that is computed using . Finally, we integrate: ( ( )) ( )

Verification Using Mathcad

To verify the above example using Mathcad we can easily find the double integral on the right hand side ( )

On the left hand side we need to specify the limits of the triple integral. Since the integral is over the unit sphere which can be written as

So Next we can find , so the integral will be


The only point here is that the final answer should have its sign changed. You should know by now why is that? EX: Let = 1. To use Gauss Theorem we will have to parameterize each side of V. First, the cylinder, C of radius 1: ( ) ( ) . We integrate over the volume V which is

inside the cylinder x2 + y2 = 1, above the plane z = 0, and below the plane z

Then the bottom, B (circle of radius r, with z=0): ( ) ( ) And finally the top, T (circle of radius r, with z=1): ( ) ( ) Now we must check orientations by computing normal vectors: For the cylinder C:

This vector points out of V. For the bottom B:

This vector points up, which is into V. We will have to remember to change the sign of any integral that is computed using For the top, T: .

This vector again points up, but at the top this is pointing out of V. Gauss Theorem says

We compute each of these integrals individually: For C: ( ( )) ( )

For B: ( ( )) ( )

For T: ( ( )) ( )

Hence, Verification Using Mathcad

To verify the above example using Mathcad we can easily find the double integrals on the right hand side. On the left hand side we need to specify the limits of the triple integral. Since the integral is over the unit cylinder which can be written as So

Next we can find

, so the integral will be

HW43: Integrate cube in R3. HW44: Let

over the boundary of the unit

. Calculate the integral of

over the ball

bounded by the unit sphere. HW45: Let the plane z = 2. Calculate HW46: The surfaces C and D are defined by 1. C is the graph of the cylindrical equation . 2. D is the set of points in the plane z = 1, which are within 0.5 unit away from the point (0.5, 0, 1). Let W be the vector field . Calculate in R3, where . Let V be the region between the cylinders

of radii 1 and 2 (centered on the z-axis), in the positive octant, and below

HW47: Suppose

, and S1 and S2 are oriented surfaces with the

same oriented boundary. Show that

(For simplicity you may assume that S1 and S2 only meet in their boundary.)

You might also like