Sie sind auf Seite 1von 14

Homework: 9/15

12.1
12. Find an equation of the sphere with center (2, -6, 4) and radius 5. Describe its intersection
with each of the coordinate planes.

Equation of the sphere:


General equation for a sphere: (x – h)2 + (y – k)2 + (z – l)2 = r2; center = (h, k, l); r = radius.

Therefore: (x – 2)2 + (y – (-6))2 + (z – 4)2 = 52  (x – 2)2 + (y + 6)2 + (z – 4)2 = 25

Intersection with xy-plane:


If we imagine that we are going to explore the sphere, but stay on the xy-plane, then we know
that wherever we are, our z-coordinate will always be 0 (or else we would be off the xy-plane).
To see the effects of this, we let z = 0 in the equation above:

(x – 2)2 + (y + 6)2 + (z – 4)2 = 25


 (x – 2)2 + (y + 6)2 + (0 – 4)2 = 25
 (x – 2)2 + (y + 6)2 + 16 = 25
 (x – 2)2 + (y + 6)2 = 9

This we identify can identify as an equation for a circle (you can confirm this by graphing it on
paper). The general equation is: (x – h)2 + (y – k)2 = r2. We see that h = 2, k = -6, and r2 = 9.
Therefore, we describe this intersection as a circle that exists on the xy-plane, center (2, -6, 0),
and radius √9 = 3.

Intersection with xz-plane:


Using the same reasoning as above, we allow y = 0:

(x – 2)2 + (y + 6)2 + (z – 4)2 = 25


 (x – 2)2 + (0 + 6)2 + (z – 4)2 = 25
 (x – 2)2 + 36 + (z – 4)2 = 25
 (x – 2)2 + (z – 4)2 = -11

From here, we see that (x – 2)2 will always be positive (since it’s a real number squared), and for
the same reason, (y + 6)2 will always be positive. Therefore, we know that they cannot sum to a
negative number (-11). So, there are no points which satisfy these conditions.

So when we tried to explore the sphere, but stay on the xz-plane, it turns out we couldn’t. There
is no point in all of 3-dimensional space that satisfies our two equations (equation of the sphere
and y = 0). Thus, the intersection is null.
Intersection with yz-plane:

Using the same reasoning as above, we allow x = 0:

(x – 2)2 + (y + 6)2 + (z – 4)2 = 25


 (0 – 2)2 + (y + 6)2 + (z – 4)2 = 25
 4 + (y + 6)2 + (z – 4)2 = 25
 (y + 6)2 + (z – 4)2 = 21

Using the same reasoning as above, we have a circle that exists on the xz-plane, center (0, -6, 4),
radius √21.

14. Find an equation of the sphere that passes through the origin and whose center is (1,2,3).

General equation for a sphere: (x – h)2 + (y – k)2 + (z – l)2 = r2; center = (h, k, l); r = radius.
So if they ask us to find an equation of the sphere, we need to find out 4 things: h, k, l, and r2.

Because they give us the center, we can line up “h” with 1, “k” with 2, and “l” with 3. Therefore,
we get: (x – 1)2 + (y – 2)2 + (z – 3)2 = r2. The only thing left is to find “r” the radius.

Humor me and consider the sphere with the equation: x2 + y2 + z2 = 1. I know that (1, 0, 0) is a
viable point on this sphere, because if I plug it in, the numbers work out: (1)2 + (0)2 + (0)2 = 1.
We are told that the sphere passes through the origin; in other words, the origin is a viable point
on the sphere. So similar, if we plug it in, the numbers should work out:
((0) – 1)2 + ((0) – 2)2 + ((0) – 3)2 = r2
 (1) + (4) + (9) = r2
 14 = r2

Therefore, we have all 4 pieces of information: (x – 1)2 + (y – 2)2 + (z – 3)2 = 14

18. Show that the equation represents a sphere, and find its center and radius:
3x2 + 3y2 + 3z2 = 10 + 6y + 12z

To show that the equation is a sphere, we must get it to form the general equation of a sphere:
(x – h)2 + (y – k)2 + (z – l)2 = r2. This should be no problem as long as we’re organized:

3x2 + 3y2 + 3z2 = 10 + 6y + 12z


 3x2 + 3y2 – 6y + 3z2 – 12z = 10
: 3(x2) (complete the square for each variable)
2
 3x + 0
: 3(y2 – 2y)
 3(y2 – 2y + 1)
 3y2 – 6y + 3
: 3(z2 – 4z)
 3(z2 – 4z + 4)
 3z2 – 12z + 12
 3x2 + 0 + 3y2 – 6y + 3 + 3z2 – 12z + 12 = 10 + 0 + 3 + 12
 3(x2) + 3(y – 1)2 + 3(z – 2)2 = 25
 (x – 0)2 + (y – 1)2 + (z – 2)2 = 25/3

Then we line this equation up with the general equation to find:


center = (0, 1, 2); radius = √(25/3) = 5/√3

12.2
14. Find a vector a with representation given by the directed line segment AB. Draw AB and the
equivalent representation starting at the origin.
A(4, 0, -2), B(4, 2, 1)

What the book means by “with representation given by” is explained on page 818. But we can
also understand that since A and B are points, rather than vectors, they don’t form vectors by
themselves. They imply a vector exists between them that will get you from A to B. So, how do
we get from A to B? Well, let’s look at it by breaking down the components. To get from “A’s
x” to “B’s x” is xB – xA = 4 – 4 = 0 (NB: to find difference we always do endpoint minus
startpoint). So if we were to draw a line from A to B in 3-dimensional space, we would find that
that line (which represents a vector) would have x-dimension 0. Similarly, yB – yA = 2 – 0 = 2,
and zB – zA = 1 – (-2) = 3. Thus our vector is ‹0, 2, 3› = a.
26. Find a vector that has the same direction as ‹-2, 4, 2› but has length 6.

If you’ve ever watched the movie Despicable Me, remember “Vector”? He called himself so
because he had magnitude (or length) and direction. All vectors can be broken down in this way.
The problem asks us to use the same direction, but a different magnitude. So let’s dissect our
vector ‹-2, 4, 2› into its direction and magnitude.

To get a vector with the same direction, we just need to multiply what we have by a scalar, and it
should stay the same. After all, a scalar just stretches or shrinks what we already have. So the
real problem is getting its magnitude to 6. First let’s get its magnitude to 1 by dividing our
current vector by its current magnitude:
|‹-2, 4, 2›| = √((-2)2 + (4)2 + (2)2) = √(4 + 16 + 4) = √24 = 2√6
‹-2, 4, 2›/2√6 = ‹-1/√6, 2/√6, 1/√6›

So right now our vector has length 1 (you can check by taking its magnitude). To get it to length,
6, we multiply it by a scalar; namely, 6:
‹-1/√6, 2/√6, 1/√6› * 6 = ‹-√6, 2√6, √6›

28. What is the angle between the given vector and the positive direction of the x-axis?
8i + 6j

Theorem 3 will relate vectors’ components/magnitudes and the angle between them. It will not
give us the angle between a vector and an axis… per se. We don’t need to use the x-axis in our
equation. An angle isn’t determined by the length of the rays/segments that form it. It’s only
determined by the direction of the rays/segments. So instead of the x-axis, we can substitute in a
vector in the same direction (that is, on the x-axis). The easiest example: ‹1, 0›, though anything
of the form ‹a, 0› will work.

8i + 6j  ‹8, 6›
By Theorem 3:
‹8, 6›•‹1, 0› =|‹8, 6›||‹1, 0›|cosθ
 (8)(1) + (6)(0) = √(82 + 62) * cosθ
 8/10 = cosθ
 cos-1(4/5) = θ

Alternatively:

We can draw a picture:


From here we want to find the angle θ. So, using trigonometry:
tanθ = 6/8
 θ = tan-1(3/4)
 θ ≈ 36.9°

31. A quarterback throws a football with angle of elevation 40° and speed 60 ft/s. Find the
horizontal and vertical components of the velocity vector.

Let’s start with a picture:

where |v| = 60. So we can imagine this as a geometry problem where the hypotenuse is 60, and
we have one of the angles.
Horizontal component: cos(40°) = x/60  x = 60cos(40°) ≈ 45.96 ft/s
Vertical component: sin(40°) = y/60  y = 60sin(40°) ≈ 38.57 ft/s
41. Find the unit vectors that are parallel to the tangent line to the parabola y = x2 at the point
(2,4).

We start by finding the tangent line to the parabola y = x2. From single variable calculus, we
know that we can find the slope of this tangent line by taking the derivative of the function:
y’(2) = 2x|x = 2 = 4.

So now that we have the slope, in order to get the unit vector that’s parallel to the tangent line,
let’s first find any vector that’s parallel to the tangent line, and then normalize it. The easiest
example of a vector with slope 4 would be ‹1, 4› (NB: “rise” = 4, “run” = 1). But this does not
have magnitude 1. To normalize we first find its magnitude: |‹1, 4›| = √17. Keep in mind that
parallel does not imply what direction the vectors are facing. So, we must also consider the
direction running in the opposite direction as ‹1, 4›/√17. Therefore, the two unit vectors are:
±(1/√17)‹1, 4›

NB: We do not need to find the equation of the tangent line. Vectors are magnitude and
direction; they do not indicate where they exist, so finding the “b” (y-intercept) of y = mx + b (of
the tangent line) gives us no useful information.

12.3
24.a. u = ‹-3, 9, 6›, v = ‹4, -12, -8›
Test for parallelism: -3/4 vs. 9/-12 vs. 6/-8  -3/4 = 9/-12 = 6/-8
Therefore, parallel. If two vectors are parallel, (barring zero vectors), they are not orthogonal, so
we do not need to check for orthogonality.

24.b. u = i – j + 2k, v = 2i – j + k
Test for parallelism: 1/2 vs. -1/-1 vs 2/1  1/2 =/= -1/-1 =/= 2/1
Therefore, not parallel

Test for orthogonality: (1)(2) + (-1)(-1) + (2)(1) = 5 =/= 0


Therefore, not orthogonal

24.c. u = ‹a, b, c› v = ‹-b, a, 0›


Test for parallelism: a/-b vs. b/a vs. c/0  Not parallel

Test for orthogonality: (a)(-b) + (b)(a) + (c)(0) = 0


Therefore, orthogonal.

26. Find the values of x such that the angle between the vectors ‹2, 1, -1›, and ‹1, x, 0› is 45°.
Theorem 3 gives us a way to relate the components/magnitudes of vectors and their angle. If we
try to substitute as much information as we can into Theorem 3:
‹2, 1, -1›•‹1, x, 0› = |‹2, 1, -1›||‹1, x, 0›|cos(45°)
 2 + x + 0 = √6 * √(1 + x2) * √2/2
 2 + x = √3 * √(1 + x2)
 x2 + 4x + 4 = 3(1 + x2) (NB: We squared both sides; we must verify solutions at the end)

From here, careful algebra will yield:


2x2 – 4x – 1 = 0
From there, the quadratic equation will give us x = 1 ± √6/2

BUT we have to make sure that both solutions work. Consider the step 2 + x = √3 * √(1 + x2).
Suppose that for one of the values of x we got in the end, the right side is positive, and the left
side is negative. Then, they’re clearly not equal. But because we squared both sides, we
wouldn’t know; we turned both of them positive. Plugging back into any of the steps before we
squared both sides does verify for us that both solutions for x are valid, but we must check.

28. Let us call our unit vector u = ‹a, b› (we’ll start with just one). Theorem 3 says that u•v = |
u||v|cosθ, where θ is the angle between vectors u and v. Since v = ‹3, 4›, and we want θ = 60°,
we can substitute these values in:
3a + 4b = |a2 + b2||32 + 42|cos(60°)
 3a + 4b = (1)(5)(1/2) (keep in mind that |u| = 1 by construction)
 b = 5/8 – 3/4*a

We have narrowed down our information into one straightforward equation with two variables.
Therefore, we need one more to solve for the variables. We only have one apparent choice:
a 2 + b2 = 1 (continue to keep in mind that |u| = 1 by construction)
2 2
 a + (5/8 – 3/4*a)

With careful algebra, we arrive at the equation: 100a2 – 60a – 39 = 0. Using the quadratic
equation, we can solve for our two values of a. For each value of a, we can substitute back into
the equation b = 5/8 – 3/4*a to find the corresponding values of b. In the end:

‹(3 + 4√3)/10, (4 – 3√3)/10› ≈ ‹0.9928, -0.1196›


‹(3 – 4√3)/10, (4 + 3√3)/10› ≈ ‹-0.3928, 0.9196›

Alternatively:
If we take the x-axis as a reference point, our two unit vectors will be angle α and β displaced
from the x-axis. We can solve for both of them since α = 60° + θ, β = 60° – θ, and we can find θ:

Trigonometry will yield that tanθ = 4/3  θ = tan-1(4/3) = 53.13°.

Then α = 60° + 53.13° = 113.13° and β = 60° – 53.13° = 6.87°

So if we have the direction of our vectors, and the magnitude (1, because they’re unit vectors),
then we can easily find the components:
u = ‹cos(113.13°), sin(113.13°)›
= ‹-0.3928, 0.9196›
v = ‹cos(-6.87°), sin(-6.87°)› (β is below the x-axis, and so we use the negative form)
= ‹0.9928, -0,1196›

(Credit for this method goes to a student in section 3)

32. Before we find the angle which they make, let us first try and establish information about
how they intersect. Where two functions intersect, their y-coordinate and x-coordinate must be
equal, so we say that y = sin(x) and y = cos(x)  sin(x) = cos(x). With the domain 0 ≤ x ≤ π/2,
this occurs only when x = π/4, where both functions equal √2/2.

Let us now try and find the slope of the functions at this point. By applying a derivative:
(sin(x))’|π/4 = cos(x)|π/4 = √2/2
(cos(x))’|π/4 = -sin(x)|π/4 = -√2/2

One way to go from here is to find out what those two vectors (tangent to our two functions) are
in terms of functions, and then find their angle. But what do we really want? We really just
wanna find out that angle θ. So we technically just have to make sure we reproduce the slopes of
those two lines. The clear candidate for slope √2/2 is ‹1, √2/2› (notice, the “rise” is √2/2, and the
“run” is 1, so slope is indeed √2/2), and for slope -√2/2, is ‹1, -√2/2›.

Then we use Theorem 3, as that will almost always be our go-to formula for finding out angles:
‹1, √2/2›•‹1, -√2/2› = |‹1, √2/2›||‹1, -√2/2›|cosθ
 (1)(1) + (√2/2)(-√2/2) = (√(1)2 + (√2/2)2)(√(1)2 + (-√2/2)2)cosθ
 1/2 = 3/2 * cosθ
 cos-1(1/3) = θ
45. Show that the vector orthab = b – projab is orthogonal to a. (It is called an orthogonal
projection of b.)

We can prove orthogonality of two vectors a and b by taking the dot product and verifying that it
equals 0. So let’s do that:

orthab • a
 (b – projab) • a
 b • a – (projab) • a (Distribute the dot product)
 b • a – ((a • b / |a|2)a) • a (projab = (a • b / |a|2)a)
 b • a – (a • b / |a|2)* a • a
 b • a – (a • b / |a|2)*|a|2 (a • a = |a|2)
 b • a – (a • b) = 0 (b • a = a • b)

47. If a = ‹3, 0, -1›, find a vector b such that compab = 2.

Let’s agree to represent b as ‹b1, b2, b3›


By definition, compab = a • b / |a|. So:
a • b / |a| = 2
 a • b = 2|a|
 (3)b1 + (0)b2 + (-1)b3 = 2√10

So we’re left with the equation 3b1 – b3 = 2√10. This might bother us if they asked us to “find
the vector b such that…” but they asked us to “find a vector b such that…”, so we can set b1, b2,
and b3 to whatever we want as long as we follow that one equation. The simplest one would be
‹0, 0, -2√10›. As long as you check your solution to make sure it fits, it should be okay.

61. Use Theorem 3 to prove the Cauchy-Schwartz Inequality:

|a • b| ≤ |a| |b|

This is arguably the most important inequality in all of analysis, so familiarizing yourself with it
(namely, being able to prove it) will be very useful.

We start with their hint: Theorem 3 says a • b = |a||b|cosθ, where θ is the angle between the two
vectors a and b. To find out where to go from here, we do a “Spot-the-Difference”:

(1) |a • b| ≤ |a| |b|


(2) a • b = |a| |b|cosθ

I count three differences:


1) the left side in (2) is not in “absolute values”.
2) the “≤” in (1) is a “=” in (2)
3) (2) has a “cosθ”

We know that a • b = |a||b|cosθ is true (i.e. equation (2)), so we can start from there and try to get
to equation (1). Applying 2) or 3), you’ll find, is difficult to do without making the equation
uglier. Let’s apply 1) then, since that’s a clean modification:

a • b = |a||b|cosθ
 |a • b| = | |a| |b| cosθ |
 |a • b| = |a||b| |cosθ| (3)

We seem to be on the right track. The problem still exists that differences 2) and 3) are still
there. But the trick to this proof is to realize that |cosθ| ≤ 1 for any real value θ.***

|cosθ| ≤ 1
 |a||b| |cosθ| ≤ |a||b| (4) (multiply both sides by |a||b|)

Well then let’s combine (3) and (4):


|a • b| = |a||b| |cosθ| ≤ |a||b|

Getting rid of the middle man:


|a • b| ≤ |a| |b|

And we’re done.

Fast track:
a • b = |a||b|cosθ (Theorem 3)
 |a • b| = | |a| |b| cosθ| (Property of “absolute values”)
 |a • b| = |a||b| |cosθ| (Property of “absolute values”)
~~~
|cosθ| ≤ 1 (Properties of cosθ)
 |a||b| |cosθ| ≤ |a||b| (multiply both sides by |a||b|)
~~~
|a • b| = |a||b| |cosθ| ≤ |a||b| (Substitution)
 |a • b| ≤ |a| |b| (Simplify)

***But Eddie, how could anyone expect someone to just guess that that’s important to the proof?

Very good question. The way to encourage those sorts of revelations is to really identify what
you want, and what you have; in that order. |a • b| ≤ |a| |b| is what we want, |a • b| = |a||b| |cosθ|
is what we have. Why does that help us? Well realize that if we can’t make |a • b| less than or
equal to |a||b|, how about we make |a||b| |cosθ| less than or equal to |a||b|?
Does this idea hold up to scrutiny though? That is, is |a||b| |cosθ| ≤ |a||b|?
At this point you can maybe see the fact that |cosθ| ≤ 1 from dividing both sides by |a||b|. But be
very careful. It would be completely incorrect and nullify your proof if you put down as one of
your steps “|a||b| |cosθ| ≤ |a||b|” before noting that |cosθ| ≤ 1. Usually when writing a proof, you
will have to work from what you want and see if you can get down to a fact that you already
have. But proofs work in the reverse order. That’s why it might seem like some proofs just
magically know why some steps will work.

If you can’t see that |cosθ| ≤ 1, try approaching it analytically. The two sides of the equation are
identical aside from that cosθ value. Well, could |a||b| |cosθ| ever be greater than |a||b|? If we
imagine |cosθ| is just a scalar, then it would have to be greater than 1 to make that happen. But
with our knowledge of cosθ, we know that can’t happen. Why? Hmm… well I guess because |
cosθ| ≤ 1. Tada!

62. The Triangle Inequality for vectors is

|a + b| ≤ |a| + |b|

(a) Give a geometric interpretation of the Triangle Inequality.

Well, if the above is the vector/analytical interpretation of the Triangle Inequality, then the
geometric interpretation (in “2-dimensional” terms) starts with recognizing that the right side is
the length of a plus the length of b. The left side is the length of a plus b, but looking at it
strictly geometrically, this is the vector composed of a and b; or rather, the “third side” (refer to
diagram below). Then we get the middle-school statement “The sum of the lengths of any two
sides is greater than or equal to the length of the third side”.
A slightly more complicated geometric interpretation is “The sum of the lengths of any two
vectors is greater than or equal to the length of their resultant (sum) vector”.

(b) Use the Cauchy-Schwarz Inequality from Exercise 61 to prove the Triangle Inequality.
[Hint: Use the fact that |a + b|2 = (a + b) • (a + b) and use Property 3 of the dot
product.]

Waste not, want not~


Cauchy-Schwarz Inequality: |a • b| ≤ |a| |b| (1)
Property 3 of dot product: a • (b + c) = a • b + a • c (2)
Other one: |a + b|2 = (a + b) • (a + b) (3)

We want: |a + b| ≤ |a| + |b|

Let’s start with what we want only for scratch work:


|a + b| ≤ |a| + |b| (4)
Of what we have, (3) seems to be the nearest to being applicable. The difference on the left sides
is that (3) is squared, while (4) is not. So let’s square both sides:

|a + b| ≤ |a| + |b|
 |a + b|2 ≤ (|a| + |b|)2 (5)

BUT squaring both sides always makes equations, and especially inequalities, more complicated.
If one side was negative, and the other was positive, because we squared both sides, we wouldn’t
know; we turned both of them positive. So we have to verify that the left side and the right side,
before squaring, are both positive (or negative, but then when we square both sides, we’d flip the
inequality (verify this for yourself if you’re not convinced)). This is simple enough, since the
left side is in “absolute values”, which is always positive. The right side is the sum of two
“absolute values” so it must also be positive. There; all justified. Continuing:

 |a + b|2 ≤ |a|2 + 2|a||b| + |b|2


 (a + b) • (a + b) ≤ |a|2 + 2|a||b| + |b|2 (We can now apply (3); our original goal)

From here it seems (2) is our next applicable step: (a + b) • (a + b) = (a + b) • a + (a + b) • b.


 (a + b) • a + (a + b) • b ≤ |a|2 + 2|a||b| + |b|2 (From equation (2))
2 2
 a • a + 2(a • b) + b • b ≤ |a| + 2|a||b| + |b| (Also from equation (2))
2 2 2 2
 |a| + 2(a • b) + |b| ≤ |a| + 2|a||b| + |b| (a • a = |a|2)
 2(a • b) ≤ 2|a||b| (Simplify)
 a • b ≤ |a||b| (Simplify)

This is really close to the Cauchy-Schwarz Inequality. But consider that if |a • b| is always less
than or equal to |a||b|, a • b, which can even be negative, should definitely be less than |a||b|.

 |a • b| ≤ |a||b|
That’s great. Now for the actual proof, we start from here, and then get to what we want.

We would be almost done, except remember that we squared both sides during our scratch work?
Well if we’re going to go backwards, then we’re going to end up taking a square root, and taking
a square root is just as dangerous in a proof as squaring both sides. We have to justify that taking
the square root won’t get messed up by negative values. Because both sides of (5) are being
squared, it’s reasonable to conclude that we’re not taking the square root of anything negative.
But what if the stuff on the left side is positive, and the stuff on the right is negative? For
example, (5)2 ≤ ((-3) + (-5))2 is a true statement, but (5) ≤ (-3) + (-5) is not. But this is also easily
verifiable, since everything inside the 2 on the left is in “absolute values”, and on the right side,
it’s the sum of two “absolute values”. Thus, we can take the square root without fear.

Fast track:
|a • b| ≤ |a||b| (From equation (1))
a • b ≤ |a||b|
2(a • b) ≤ 2|a||b|
|a|2 + 2(a • b) + |b|2 ≤ |a|2 + 2|a||b| + |b|2
a • a + 2(a • b) + b • b ≤ |a|2 + 2|a||b| + |b|2 (a • a = |a|2)
(a + b) • a + (a + b) • b ≤ |a|2 + 2|a||b| + |b|2 (From equation (2))
(a + b) • (a + b) ≤ |a|2 + 2|a||b| + |b|2 (Also from equation (2))
|a + b|2 ≤ |a|2 + 2|a||b| + |b|2 (From equation (3))
|a + b|2 ≤ (|a| + |b|)2

Since the left side and the right side of the inequality are either nonnegative by definition or the
sum of values nonnegative by definition, we if square root both sides the inequality holds.

|a + b| ≤ |a| + |b| (Square root both sides)

Q.E.D.

(Disclaimer: this solution set is subject to typos or careless errors. I apologize in advance for
any of those.)

Das könnte Ihnen auch gefallen