Math GRE Practice Solutions
Math GRE Practice Solutions
UNOFFICIAL TESTS
SOLUTIONS
Answers
1. B 23. A 45. A
2. A 24. D 46. E
3. A 25. D 47. E
4. E 26. C 48. A
5. D 27. C 49. D
6. C 28. E 50. A
7. C 29. C 51. E
8. B 30. D 52. D
9. D 31. D 53. E
10. D 32. D 54. C
11. A 33. E 55. A
12. B 34. A 56. C
13. B 35. A 57. C
14. B 36. C 58. B
15. A 37. E 59. B
16. B 38. A 60. D
17. E 39. E 61. C
18. D 40. C 62. A
19. B 41. B 63. E
20. C 42. D 64. A
21. C 43. A 65. C
22. A 44. D 66. B
FORM MSDC01
MATHEMATICS TEST
SOLUTIONS
GRE is a registered trademark of Educational Testing Service, which does not sponsor or endorse this product.
Z 2 p
1. 16 − 4x2 dx =
−2
1
2. Four semicircular arcs are inscribed in a square as shown in the figure above. Find the ratio
of the shaded area to the area of the square.
The shaded area is the difference of a quarter-circle with area 14 πx2 and a right triangle with
area 21 x2 , giving us ( 14 π − 21 )x2 . There are eight of these figures in the overall picture, so the
shaded area is (2π − 4)x2 . Thus the ratio of the shaded area to the area of the square comes
(2π − 4)x2 1
out to 2
= (π − 2).
4x 2
2
3. The line y = x + 1 is tangent to which of the following curves at x = 1?
√
(A) y = x
√
(B) y = x+1
√
(C) y = x−1
√
(D) y =2 x
√
(E) y =2 x−1
Solution:
When x = 1, we have y = 1 + 1 = 2, so the curve must pass through (1, 2); eliminate A, C,
√ dy 1
and E. Then, the derivative of y = x + 1 is = √ , which evaluated at x = 1 gives us 21 .
dx 2 x
√ dy 1
On the other hand, the derivative of y = 2 x is = √ , which evaluated at x = 1 gives us
dx x
1. Since the line y = x + 1 has a slope of 1, the answer must be D.
3
4. What are the most specific conditions under which the statement P ∧ (P → Q) → Q is
true?
4
5. Suppose f and g are continuously differentiable functions with the following properties:
f (x) > 0 and g(x) > 0 for all x ∈ R.
f 0 (x) > 0 for all x ∈ R.
g 0 (x) < 0 for all x ∈ R.
Which of the following functions is NOT necessarily monotonic?
2 f (x)
(A) g(x) (B) f (x) − g(x) (C) f (x)g(x) (D) (E) g f (x)
g(x)
Solution:
Let’s work out each in turn by differentiation, looking at the signs we get:
d 2
A: g(x) = 2g(x)g 0 (x) = 2(+)(−) < 0
dx
d
f (x) − g(x) = f 0 (x) − g 0 (x) = (+) − (−) > 0
B:
dx
d
C: f (x)g(x) = f 0 (x)g(x) + f (x)g 0 (x) = (+)(+) + (+)(−)
dx
d f (x) g(x)f 0 (x) − f (x)g 0 (x) (+)(+) − (+)(−)
D: = 2 = >0
dx g(x) g(x) (+)2
d
g f (x) = g 0 f (x) f 0 (x) = g 0 (+) · (+) = (−)(+) < 0
E:
dx
Notice that with option C, f 0 (x)g(x) is positive while f (x)g 0 (x) is negative; however, we have
no idea which is greater magnitude, so we cannot conclude the sign for any given x. It’s
entirely possible for this result to be positive for some x and negative for others, so f (x)g(x)
does not have to be monotonic.
5
6. Let f : R → R be a function defined on the real numbers. Which of the following ensures
that lim f (x) = ∞?
x→−∞
(A) For all ε < 0, there exists a δ < 0 such that x < δ implies f (x) < ε.
(B) For all ε > 0, there exists a δ < 0 such that x > δ implies f (x) > ε.
(C) For all ε > 0, there exists a δ < 0 such that x < δ implies f (x) > ε.
(D) For all δ > 0, there exists an ε < 0 such that x > δ implies f (x) > ε.
(E) For all δ < 0, there exists an ε > 0 such that x < δ implies f (x) > ε.
Solution:
This one is really about just remembering how limits are defined when involving infinity.
Since x needs to approach −∞, it needs to be able to be less than any negative real number
δ, since f (x) needs to approach ∞, it needs to be able to be greater than any positive real
number ε. Past that, just remember every analysis student’s favorite phrase: “For every ε
there exists a δ...”
6
z
7. Suppose z is a complex number such that |z| = 8. Which of the following is equal to ?
4
(Here z stands for the complex conjugate of z.)
z z 1 16 16
(A) (B) (C) (D) (E)
16 16 16z z z
Solution:
√
Recall that |z| = z · z, so we can calculate hence:
√
z·z =8
z · z = 64
64
z=
z
z 1 64 16
= · =
4 4 z z
If you didn’t remember the above equation for |z|, though, you could always try letting z = 8i
and eliminating the other choices! (You wouldn’t want to try z = 8 because then you can’t
tell the difference between z and z.)
7
8. For which of the following shapes does the set of rotation and reflection symmetries form
an abelian group? (Assume that sides and angles that appear to be congruent are in fact
congruent.)
Solution:
Let Dn stand for the dihedral group of order 2n, which consists of n reflection symmetries
and n rotation symmetries 360/n degrees apart, including the identity transformation.
For n = 3 or greater, Dn is NOT abelian, so this eliminates B (D3 ), C (D5 ), and D (D4 ).
However, E only has two lines of reflection symmetry (one horizontal and one vertical) and
two rotations (0◦ and 180◦ ), so its group is D2 . This is isomorphic to the Klein 4-group V4 ,
which is abelian.
The circle, which has infinitely many reflection and rotation symmetries, has as its symmetry
group the orthogonal group O(n, R), but you don’t really need to know that — all you need
to know is that reflections and rotations (other than 0◦ or 180◦ ) don’t commute in general.
8
Z x2 √
9. Let g(x) = cos( t) dt. What is the value of g 0 (π)?
3
9
10. Let P = (3, −1, 2, 5) and Q = (1, 2, 2, 4) be two points in R4 . Which of the following points
lies on the line in R4 connecting P and Q?
(A) (5, 5, 2, 3)
(B) (1, 5, 3, −1)
(C) (−1, 3, 2, 0)
(D) (−2, 6, 3, 1)
(E) (−3, 8, 2, 2)
Solution:
The line connecting P and Q can be parametrized as
This instantly eliminates B and D since their third coordinates are not 2. From here, we look
at each remaining option in turn:
• If 3 − 2t = 5, then t = −1, but this would imply that −1 + 3(−1) = 5. Eliminate A.
• If 3 − 2t = −1, then t = 2, but this would imply that −1 + 3(2) = 3. Eliminate C.
• If 3 − 2t = −3, then t = 3, which also implies that −1 + 3(3) = 8 and 5 − 3 = 2.
10
11. If f (x) = x1/x , what is f 0 (2)?
√ √ √ √ √
2 2 2 2 2
(A) + log 2 (B) log 2 (C) (1 − log 2) (D) (1 + log 2) (E) (−1 + log 2)
4 4 4 4 4
Solution:
The standard way to work this is to use logarithmic differentiation:
y = x1/x
log y = log x1/x
1
= log x
x
d d 1
log y = log x
dx dx x
y0 1 1 −1
= · + 2 log x
y x x x
1
y 0 = y · 2 (1 − log x)
x
√
1/2 1 2
Substituting x = 2, we have x = 2 · 2 (1 − log 2) = (1 − log 2).
2 4
There’s a neat trick that can make this quicker though. It turns out that if f and g are both
functions of x, then while f g is neither a power function nor an exponential function, we can
treat it first like a power function and then like an exponential function, and add the two
results together:
d g
f = gf g−1 · f 0 + f g ln f · g 0
dx
11
1
12. Brian is playing a crane game where the chance of winning a plush toy is each time. What
3
is the probability that it takes him exactly 5 tries to win 3 plush toys?
2 8 4 10 40
(A) (B) (C) (D) (E)
81 81 243 243 243
Solution:
2 3
2 1 4
The probability of losing two games and winning three, in that order, is · = 5.
3 3 3
However, there are multiple orders in which these games could have happened; the fifth
game is always a win, but the first four games consist of two wins and two losses. Thus we
need the number of arrangements of the “word” WWLL, which is
4! 24
= = 6.
2!2! 2·2
4 8
Multiplying the above probability by this number of arrangements, we have 6 · 5
= .
3 81
12
13. A supplier is manufacturing disposable paper cups in the shape of a right circular cone. The
slant height of each cone is to be 4 inches long. What should be the diameter of the open
circular base of the cup, so that it holds the maximum possible volume of water?
√ √ √ √ √
4 3 8 3 2 6 4 6 8 6
(A) (B) (C) (D) (E)
3 3 3 3 3
Solution:
1 2
First let’s sketch the situation: The volume of the cone is V = πr h. Since the slant height
3
is 4, we know that r2 + h2 = 42 .
At this point, we could either eliminate r or h; however, eliminating r makes the algebra
much nicer, so that we don’t have to deal with square roots! Let r2 = 16 − h2 , which gives us
1 16 1
V = π(16 − h2 )h = πh − πh3 .
3 3 3
r
dV 16 2 16
Differentiating, we have = π − πh ; solving for h, we have h = . This means that
r dh
√ 3 3
16 32 4 6
r = 16 − = = . Therefore the diameter that maximizes the volume of the paper
3
√ 3 3
8 6
cup is d = inches.
3
13
Z x p
14. For how many values of x does t 9 + t2 dt = 0?
−3
(A) None (B) One (C) Two (D) Three (E) Four
Solution:
First of all, if we let x = −3, then our integral covers no area and we get 0.
√
Next, suppose x = 3. Since t is an odd function and 9 + t2 is an even function, their product
is odd. This means we’re integrating an odd function over a symmetric interval, so we again
get 0.
To see
Z xthat these are the only solutions, notice that by the Fundamental Theorem of Calculus,
d p p
t 9 + t2 = x 9 + x2 , which only is zero at t = 0. This means that the function
dx −3
defined by our above integral only changes direction once, so the two zeros we found are the
only ones.
14
15. Which of the following most closely resembles the graph of the curve defined by the polar
equation r = 1 − 2 cos θ?
(D) (E)
Solution:
First of all, notice that since r is a function of cos θ, we have
Therefore the curve should be symmetric about θ = 0, which is the horizontal axis. Eliminate
A and D.
Now notice that the curve should touch the origin whenever r = 0. Setting 1 − 2 cos θ = 0, we
1 π π
get cos θ = , which occurs at θ = and θ = − . The only remaining curve that intersects
2 3 3
the origin twice is E.
15
16. For a hexadecimal (base 16) number, let the digits a, · · · , f represent the numbers 10, · · · , 15
in base 10. Which of the following is a divisor of 14dhex ?
(A) 1chex (B) 25hex (C) 28hex (D) 2dhex (E) 33hex
Solution:
As stated in the directions, the hexadecimal number system has the digits 0-9, as well as
the extra digits a, b, c, d, e, and f, which correspond to 10, 11, 12, 13, 14, and 15 in base 10
respectively.
First let’s convert 14d to base 10:
16
1 −2 0 2
17. Let A = 3 2 −1 4. Which of the following is a basis for the null space of A?
1 6 −1 0
1 −1
(A) 3 , 1
1 3
0
0
(B)
0
0
−4
2
(C)
8
4
2
−6
1 1
(D) ,
8 0
0 4
−4 2 −6
, , 1
2 1
(E) 8 8 0
4 0 4
Solution:
The null space of a matrix A is the set of all vectors v such that Av = 0. Normally to compute
this, we’d augment by a column of zeros and row-reduce, but since this is multiple-choice,
there’s no reason to do that!
A: This answer is bogus — a 3 × 4 matrix can’t be multiplied by a 3 × 1 column vector. (In
fact, this is the basis of the column space.)
B: Multiplying by the zero vector trivially gives us 0, so no surprises here. This could be
the answer if the matrix has a full rank of 3, so let’s keep going.
C: Multiplying A by this vector does in fact give us 0, so we can eliminate B.
D: Multiplying A by either of these vectors also gives us 0, and neither is a multiple of the
other. Therefore our null space has dimension at least 2; eliminate C.
E: Multiplying A by each of these three vectors gives us 0; however, notice that the first
vector is the sum of the other two. This means the set is not linearly independent, and
therefore can’t be a basis. Eliminate E, leaving D as the correct answer.
17
18. Which quadrants are contained in the preimage of quadrant III of the complex plane under
the mapping z 7→ z 3 ?
18
x 0 2 4 6
f 0 (x) 4 1 7 −1
f 00 (x) −1 3 0 −3
19. The function f is twice differentiable for all real x. Values of f 0 (x) and f 00 (x) are given for
selected values of x in the table above. Which of the following statements must be true?
I. f 0 has a local maximum at x = 4.
II. f has a point of inflection somewhere in the interval (0, 2).
III. There exists a c ∈ [0, 6] for which f 00 (c) = −4.
(A) I only (B) II only (C) III only (D) I and II only (E) II and III only
Solution:
I. f 0 has a local maximum when f 00 changes sign from positive to negative.
It’s tempting to pick this one, since for x = 2 → 4 → 6, f 00 (x) goes positive → zero →
negative. However, it’s possible that the graph of f 00 merely touches the x-axis at x = 4,
goes back up, and then crosses through it some time afterward, say at x = 5. All we
know is that there has to be a sign change somewhere between 4 and 6.
II. f has a point of inflection when f 00 changes sign.
At first we might think to use the Intermediate Value Theorem to say that there has to
be a sign change between −1 and 3. Note that we did not say that f is twice continuously
differentiable. Nevertheless, Darboux’s theorem says that the derivative of a function
still satisfies the intermediate value property, so our sign change is still guaranteed.
III. Again the Intermediate Value Theorem won’t really help us here, since the table only
shows values of f 00 between −3 and 3. However, notice that on the interval [4, 6], the
−1 − 7
average rate of change of f 0 (x) is = −4. Since f 00 is twice differentiable, f 0 is
6−4
differentiable (as well as continuous), so by the Mean Value Theorem, there must exist
a c ∈ [4, 6] such that f 00 (c) = −4.
19
20. Find the volume of the solid formed by revolving about the y-axis the region in the first
2
quadrant bounded by the curves y = e−x and x = 2.
(A) π(1 − e−4 ) (B) 2π(1 − e−4 ) (C) π(1 + e−2 ) (D) 2π(1 − e−2 ) (E) 2π(1 + e−2 )
Solution:
We can use the Shell Method to find the volume:
Z 2 Z 2
2
V = 2πxf (x) dx = 2πxe−x dx
0 0
20
21. Suppose a curve C is parametrized by the equations x = f (t) and y = g(t), where f and g
are twice-differentiable functions. If f 0 (t) 6= 0, which of the following expressions gives the
d2 y
value of 2 when it exists?
dx
Solution:
dy g 0 (t)
First, the first derivative is = 0 .
dx f (t)
d2 y dy 0
The easiest way to remember the second derivative for parametric equations is = :
dx2 dx
d
g 0 (t)
f 0 (t)g 00 (t) − g 0 (t)f 00 (t)
2
dy 0 dt f 0 (t) f 0 (t) f 0 (t)g 00 (t) − g 0 (t)f 00 (t)
= 0
= =
f 0 (t)
3
dx f (t) f 0 (t)
21
ZZ
22. Evaluate cos(x2 ) dx dy, where Ω is the triangular region pictured above.
Ω
1 1 1
(A) sin 1 (B) cos 1 (C) sin 1 (D) (1 − sin 1) (E) (1 − cos 1)
2 2 2
Solution:
In order to compute this integral, we need to switch the order of integration:
ZZ Z 1Z x
2
cos(x ) dx dy = cos(x2 ) dy dx
Ω 0 0
Z 1 x
2
= y cos(x ) dx
0 0
Z 1
= x cos(x2 ) dx
0
1 1
= sin(x2 )
2 0
1
= sin 1
2
22
ex + e−x
23. If f (x) = , then f −1 (x) =
ex − e−x
r r r r r
x+1 x−1 x+1 x−1 x+1
(A) log (B) log (C) log (D) log (E) log
2 2 x−1 x+1 1−x
Solution:
We find f −1 (x) by writing y = f (x), swapping x and y, and then solving for y again.
ex + e−x
y=
ex − e−x
ey + e−y
x= y
e − e−y
x(e − e ) = ey + e−y
y −y
At this point, we can multiply both sides of the equation by ey , which allows us to solve for
y:
xe2y − x = e2y + 1
xe2y − e2y = x + 1
(x − 1)e2y = x + 1
x+1
e2y =
x−1
x+1
2y = log
x−1
1 x+1
y = log
2 x−1
r
x+1
= log
x−1
23
24. Which of these rings has the largest number of units?
24
25. Let y = f (x) be a nonzero solution to the differential equation
y 00 + 4y 0 + 7y = 0.
(A) I only (B) II only (C) I and II only (D) II and III only (E) I, II, and III
Solution:
The characteristic equation of this differential equation is λ2 + 4λ + 7 = 0, which we can solve
with the quadratic formula:
√
√
p
−4 ± (−4)2 − 4(1)(7) −4 ± −12
λ= = = −2 ± i 3
2(1) 2
Since our solutions are nonreal, the general solution to our system of equations is
√ √
y = c1 e−2t cos( 3t) + c2 e−2t sin( 3t).
I. Since our solution oscillates around 0, there are indeed infinitely many solutions.
II. The e−2t factor means that this is a decaying oscillator, so it indeed tends toward zero.
We could prove this if we wanted using the Squeeze Theorem.
III. While the magnitude of f (x) does grow greater as x → −∞, it also returns back to 0
infinitely often, which means that this limit does not exist.
25
26. In a room with ten people, some pairs of people shake hands, while some don’t. Nobody
shakes the same person’s hand more than once. Each person in the room is then asked
whether they shook hands with an even or odd number of people. Which of the following
statements is true?
26
27. Let g be the function defined by g(x, y, z) = z 2 exy for all real x, y, and z. The maximum
possible value M of the directional derivative of g at the point (2, 0, −1) in the direction of
some vector u ∈ R3 falls within which of the following ranges?
(A) 1 < M < 2 (B) 2 < M < 3 (C) 3 < M < 4 (D) 4 < M < 5 (E) M > 5
Solution:
Recall that the directional derivative in the direction of a unit vector u is given by ∇g · u. The
maximum possible value is then the gradient ∇g:
∂g ∂g ∂g
= yz 2 exy , xz 2 exy , 2zexy
∇g(x, y, z) = , ,
∂x ∂y ∂z
√
Thus ∇g(2, 0, −1) = (0, 2, −2), the magnitude of which is 2 2 ≈ 2.828.
27
dn
kπ
28. cos x = sin x + if and only if which of the following congruences holds?
dxn 2
d kπ kπ
From here, if n = 1, then we have cos x = − sin x = sin x cos + cos x sin . This means
dx 2 2
kπ kπ
we need cos = −1 and sin = 0, which is true if k = 2. Thus k − n = 1 ≡ 1 (mod 4).
2 2
d3
kπ
Another possibility would be to let n = 3, giving us 3 cos x = sin x = sin x + . Since
dx 2
sin x has period 2π, k can be any multiple of 4, and the result follows.
28
29. Find the remainder when 6293 is divided by 11.
62 = 36 ≡ 3 (mod 11)
3 2
6 = 6 · 6 ≡ 6 · 3 = 18 ≡ 7 (mod 11)
29
sin x − tan x
30. lim =
x→0 x3
1 1
(A) −1 (B) − (C) 0 (D) (E) The limit does not exist.
2 2
Solution:
0
Seeing the indeterminate form , we could use L’Hôpital’s Rule, but if we do so in its current
0
state, the results will be hairy. Instead, noticing sin x in the numerator and at least one x in
the denominator, let’s play with the expression algebraically:
sin x
sin x − tan x sin x − cos x sin x(1 − sec x) sin x 1 − sec x
3
= = = ·
x x3 x3 x x2
sin x 1 − sec x
This means that we can split our limit into lim · lim . The former limit is 1,
x x→0 x→0 x2
which you can check with L’Hôpital’s Rule if you wish, so we just need to evaluate the latter.
Since sec 0 = 1, let’s go ahead and use L’Hôpital’s Rule:
30
31. Suppose the power series a0 + a1 (x + 1) + a2 (x + 1)2 + a3 (x + 1)3 + · · · is used to represent
2x
the function f (x) = 2 . What is the radius of convergence of this power series?
x +1
√ √
(A) 1 (B) 2 (C) 3 (D) 2 (E) ∞
Solution:
We could go through the work of finding a Taylor series expansion about x = −1 and then
determining the radius of convergence using the ratio test. However, the easiest way to think
about this is to imagine f as a function of a complex number:
2z
f (z) =
z2 + 1
Note that this function has singularities (simple poles, in fact) at z = i and z = −i. Therefore,
the disk of convergence centered at z = −1 can only extend until it hits either of these
singularities (which it actually hits both of at the same time).
√
Thus the radius of convergence is the distance from z = −1 to z = i, which is 2.
31
32. What is the set of all vectors v that satisfy the equation (2, 3, 4) × v = (−3, 2, 1)?
(A) ∅
n o
(B) (−5, −14, 13), (8, 12, −13)
n o
(C) (5, 14, −13), (−8, −6, −12)
n o
(D) (−8, −12, 13), (8, 6, −12)
n o
(E) (x, y, z) : 3x − 2y − z = 0
Solution:
Remember that for two vectors a and b, a×b will be perpendicular to both a and b. However,
notice that (2, 3, 4) · (−3, 2, 1) = (2)(−3) + (3)(2) + (4)(1) = 4. Since this dot product isn’t
zero, there’s no way that (2, 3, 4) × v could possibly be (−3, 2, 1) no matter what v is.
32
33. Suppose x is the smallest positive integer that satisfies the following congruences:
x≡1 (mod 4)
x≡2 (mod 5)
x≡3 (mod 7)
(A) x ≡ 2 (mod 9)
(B) x ≡ 4 (mod 9)
(C) x ≡ 6 (mod 9)
(D) x ≡ 8 (mod 9)
(E) There is no such value of x.
Solution:
The Chinese Remainder Theorem says that since the moduli are coprime, we can definitely
find a solution:
33
34. Consider the following algorithm, which takes an input integer n>1 and prints a decimal
number.
input(n)
t := 0
k := 1
s := 1
while (k < n) {
a := 1/k
t := t + s*a
k := k + 2
s := s * -1
}
output(t)
If the input integer is 500, which of the following will be the output when truncated after the
hundredths digit?
(A) 0.59 (B) 0.69 (C) 0.72 (D) 0.78 (E) 0.81
Solution:
The effect of this algorithm is to approximate the infinite series:
1 1 1
1− + − + ···
3 5 7
In particular:
• k is the index of the term being computed, which increases by 2 each time.
• n is the maximum index added to the summation.
• s is the sign of the term, which is repeatedly flipped by multiplying by -1.
• a is the absolute value of the term, which is then multiplied by s to produce the alternating
series.
• t is the total of all the terms so far.
Since the value of n given is 500, the last term appended is -1/499. Then the error for the
series is bounded above by the absolute value of the next term, which would have been 1/501,
guaranteeing that our approximation is correct to the nearest hundredth.
The value of this series can be found by remembering the Taylor series for arctan x:
x3 x5 x7
arctan x = x − + − + ···
3 5 7
π
The value of this series evaluated at x = 1 is ≈ 0.78.
4
34
35. Let y = f (x) be a solution to the differential equation y 0 = y 3 − 3y + 2, y(0) = a, where a ∈ Z.
For how many values of a is lim f (x) finite?
x→∞
(A) One (B) Two (C) Three (D) Four (E) More than four
Solution:
What we’re looking for is the equilibrium solutions to the differential equation, which are
found when y 0 = 0. By inspection, y 0 = 0 when y = 1; we can divide y 3 − 3y + 2 by y − 1 to
factor it, for example with synthetic division.
1 0 −3 2
1 1 1 −2
1 1 −2 0
Now we can classify what happens for initial conditions in these regions.
• If y(0) < −2, then y 0 < 0, and the solution will tend toward −∞.
• If −2 < y(0) < 1, then y 0 > 0, and the solution will tend toward a limit of 2.
• If y(0) > 1, then y 0 > 0, but the solution will tend toward +∞.
Thus lim f (x) is finite for the initial condition y(0) = a if a = −1, 0, 1, or 2.
x→∞
35
36. The above graph of y = fZ(x), defined for −2 ≤ x ≤ 3, consists of a semicircle and two line
x
segments. Define g(x) = f (t) dt, and let A, B, and C be defined as follows:
0
(A) A<B<C
(B) A<C=B
(C) B<A<C
(D) C<A<B
(E) C=B<A
Solution:
Let’s look at each of A, B, and C:
• For A, note that if we integrate backwards from 0, we get a positive result. In particular,
Z −2
π
g(−2) = g(x) dx = ≈ 1.57.
0 2
• For B, g has an inflection point wherever g 00 changes sign. Since g 00 = f 0 , we see from the
graph that f 0 changes sign at x = −1 (where f switches from decreasing to increasing)
and x = 1 (where f switches from increasing to decreasing), so g has two inflection
points.
• For C, g is not differentiable wherever g 0 is not defined. Since g 0 = f , we see from the
graph that f is defined for all x in the interval −2 ≤ x ≤ 3. (Don’t let the sharp corners
fool you — that’s where f , not g, fails to be differentiable!) So, C = 0.
Ordering these, we have C < A < B.
36
37. Suppose (x − λ1 )(x − λ2 )(x − λ3 ) is the characteristic polynomial of the matrix
3 1 0
A = 1 2 1 .
0 1 3
1 1 1
Calculate the quantity + + .
λ1 λ2 λ1 λ3 λ2 λ3
1 1 2 4 19
(A) (B) (C) (D) (E)
2 3 3 9 12
Solution:
λ1 + λ2 + λ3
The quantity we want can be rewritten as .
λ1 λ2 λ3
The sum of the eigenvalues is the trace of the matrix, which is 3 + 2 + 3 = 8. We could also
then find the determinant of the matrix, which is the product of the eigenvalues, but we can
actually figure out the eigenvalues directly by looking at the matrix:
• The rows have a constant sum of 4, so 4 is one of the eigenvalues (corresponding to the
vector (1, 1, 1)T ).
• By inspection,
subtracting 3I from the main diagonal will give two identical rows of
0 1 0 , so 3 is one of the eigenvalues.
• Since the trace is 8, the remaining eigenvalue must be 1.
8 2
Therefore the product of the eigenvalues is 3 · 4 · 1 = 12, so or desired quantity is = .
12 3
37
38. If G is a group of order 60, then G does not necessarily have a subgroup of order:
38
(−1)n n
39. Let {an } be the sequence defined by an = 1+ . Calculate lim sup an − lim inf an .
n n→∞ n→∞
e−1 e−1 e2 − 1
(A) 0 (B) (C) (D) (E) +∞
e e2 e
Solution:
x n
Recall that lim 1 + = ex . What we have here, then, is an interleaving of two sequences
n→∞ n
1 n −1 n
— one is 1+ , which converges to e, and the other is 1+ , which converges
n n
1 1 e2 − 1
to e−1 . Therefore lim sup an = e and lim inf an = , giving a difference of e − = .
n→∞ n→∞ e e e
39
40. Let P3 (R) be the vector space of all polynomials of degree at most 3 with real coefficients.
Consider the following subspaces of P3 (R):
U = {p ∈ P3 (R) : p(0) = 0}
V = {p ∈ P3 (R) : p(−1) = p(1) = 0}
Which of the following statements are true?
I. U ∩ V is a subspace of P3 (R).
II. U ∪ V is a subspace of P3 (R).
III. dim(U ) + dim(V ) = dim (P3 (R)).
(A) I only (B) III only (C) I and II only (D) I and III only (E) I, II, and III
Solution:
Remember that subspaces must be closed under addition and scalar multiplication.
I. An element of U ∩ V is a polynomial for which p(−1) = p(0) = p(1) = 0. If we add two
such polynomials together or multiply one by a real number, the zeros are unchanged,
so U ∩ V is a subspace.
II. This is not true. Let p(x) = x and q(x) = x2 − 1. Then p(x) + q(x) no longer has any of
these roots, so U ∪ V is not closed under addition and therefore is not a subspace.
III. Since cubic polynomials have four coefficients, P3 (R) is isomorphic to R4 , so it has
dimension 4. From here, every time we specify one of the roots, we introduce a linear
dependency among the coefficients of our polynomials. Suppose an element of P3 (R)
is written as p(x) = ax3 + bx2 + cx + d:
• If p(0) = 0, then we know that d = 0.
• If p(1) = 0, then we know that a + b + c + d = 0.
• If p(−1) = 0, then we know that −a + b − c + d = 0.
Each linear dependency reduces the dimension of our space by 1, so dim(U ) = 3 and
dim(V ) = 2. This means that our given equation is false; the correct statement would
be dim(U ) + dim(V ) − dim(U ∩ V ) = dim(P3 (R)), since dim(U ∩ V ) = 1.
40
Z
41. Let C be the semicircular path from (0, 0) to (2, 0) pictured above. Evaluate F · dr, where
C
F(x, y) = (x2 + y 2 )i + 2xyj.
8 16
(A) 2 (B) (C) 3 (D) 4 (E)
3 3
Solution:
We could parametrize the semicircle and evaluate our integral directly, but notice what happens
when we look at the cross-partials of the components of F:
∂ 2 ∂
(x + y 2 ) = 2y = (2xy)
∂y ∂x
This means that we’re taking the line integral of a gradient of some function, so the integral
result is path-independent by the Fundamental Theorem of Line Integrals. We could instead
just choose to integrate over the straight line path from (0, 0) to (2, 0), or better yet, we could
reconstruct the original function (the potential) from its gradient.
Z
1
(x2 + y 2 ) dx = x3 + xy 2 + g(y)
3
∂ 1 3
x + xy 2 = 2xy + g 0 (y)
∂y 3
1
Since g 0 (y) = 0, our potential function is x3 + xy 2 + C. Thus the value of our integral is
(2,0) 3
1 3 8
x + xy 2 = .
3 (0,0) 3
41
r
√ dy
q p
42. If y = x+ x + x + x + · · ·, then =
dx
1 1 y 2y + 1 1 − 2y
(A) (B) (C) (D) (E)
2y − 1 2y + 1 2y − 1 y y
Solution:
√
We can rewrite this expression as y = x + y, or y 2 = x + y. From here we can use implicit
differentiation:
d 2 d
(y ) = (x + y)
dx dx
dy dy
2y =1+
dx dx
dy dy
2y − =1
dx dx
dy
(2y − 1) =1
dx
dy 1
=
dx 2y − 1
42
43. Neisha has four index cards, each of which has a different set written on it, as shown above.
First, Neisha chooses a card at random, and lets A be the set written on that card. Then, she
replaces the card and shuffles the cards, chooses a second card at random, and lets B be the
set written on the second card. If F is the set of all functions with domain A and codomain
B, what is the probability that F is a countable set?
1 1 3 3 9
(A) (B) (C) (D) (E)
2 4 8 16 16
Solution:
Recall that the number of functions with domain A and codomain B is |B||A| .
The cardinalities of Z, Q, R, and {0, 1} are ℵ0 , ℵ0 , 2ℵ0 , and 2, respectively. (Also remember
that 2ℵ0 is strictly greater than ℵ0 , which is why we say R is uncountable.)
The only possible exponentiations that will lead to a cardinality of ℵ0 or less are 22 = 4
and ℵ0 2 = ℵ0 ; anything else will give an uncountable cardinality. This means that the only
combinations that give a countable set of functions are:
43
44. Consider the function f defined as
c
5/2
if x ≥ 1
f (x) = x
0 if x < 1,
44
45. Suppose S is a set of continuous functions on [3, 5] such that for each f ∈ S, the following
properties hold:
f (3) = 2
f (5) = 4
f 0 (x) > 0 for all x ∈ [3, 5]
Z 4
−1
Calculate sup f (x) dx .
f ∈S 2
From here, the supremum will be approached as the graph of f looks as much as possible
like what’s given in the figure above — close to 2 for as long as possible, then shooting up to
4 as x reaches 5. Thus the supremum of all possible areas to the left is 10.
45
Z x
46. For each integer n ≥ 0, define Pn (x) = tn e−t dt. Which of the following recurrences is
0
satisfied by Pn (x) for all n ≥ 1?
46
47. The vertex-edge graph above depicts a relation ∼ on the set S = {a, b, c, d}. For any x ∈ S
and y ∈ S, an arrow drawn from x to y on the graph signifies that x ∼ y.
What is the minimum number of additional arrows that must be drawn so that the relation
represented by the resulting vertex-edge graph is transitive?
(A) Two (B) Three (C) Four (D) Five (E) Seven
Solution:
We need to make sure that there is an arrow x → y whenever there is a path from x to y:
• a→b→a
• b→a→b
• c→a→b
• a→b→d
• c→a→b→d
Thus we need five additional arrows, as illustrated below.
47
1/n
nn
48. lim =
n→∞ n!
(A) 1 (B) e (C) e1/e (D) ee (E) The limit does not exist.
Solution:
n n n n 1/n
We can rewrite our limit as L = lim · · · ... · . Taking the logarithm of both
n→∞ 1 2 3 n
1 n n n n
sides gives us log L = lim log + log + log + . . . + log . This looks remarkably
n→∞ n 1 2 3 n
like a Riemann sum:
n
1X n
log L = lim log
n→∞ n k
k=1
Z 1
1
= log dx
0 x
Z 1
=− log x dx
0
1
= −(x log x − x) = 1
0
48
49. Michael brought 12 identical cookies to work. In how many ways can he distribute those
cookies to his four coworkers so that each coworker gets at least one cookie?
11 12 12 13 15
(A) (B) (C) (D) (E)
3 3 4 4 3
Solution:
This kind of problem is usually solved via the so-called “stars-and-bars” method, though we
have to deal with the condition that each coworker gets at least one cookie. We’ll do that by
giving one cookie to each coworker ahead of time, which leaves 8 cookies to distribute. Now
we think of those cookies as 8 “stars”, and we think of our 4 coworkers as being separated by
3 “bars”, all of which we can arrange in a line:
? ? ? ? | ? || ? ? ?
The arrangement above would mean that of the remaining 8 cookies, the four coworkers get
4, 1, 0, and 3, respectively.
11! 11
The number of arrangements of 8 “stars” and 3 “bars” is = .
8!3! 3
49
Z 1
sin t
50. Estimate dt to the nearest thousandth.
0 t
(A) 0.942
(B) 0.943
(C) 0.944
(D) 0.945
(E) 0.946
Solution:
This is what power series are made for! Begin with the Maclaurin series for sin t:
t3 t5 t7
sin t = t − + − + ···
3! 5! 7!
Divide through by t:
sin t t2 t4 t6
= 1 − + − + ···
t 3! 5! 7!
Integrate:
x
t3 t5 t7
Z
sin t
dt = t − + − + ···
0 t 3 · 3! 5 · 5! 7 · 7!
Evaluate at x = 1:
Z 1
sin t 1 1 1
dt = 1 − + − + ···
0 t 3 · 3! 5 · 5! 7 · 7!
1 1 1
=1− + − + ···
18 600 7 · 7!
Since 7! = 5040, 7 · 7! will be well above the error bound of 0.0005 we need to be correct to the
1
nearest thousandth, so we can just sum the first three terms. In particular, = 0.11111 · · · ,
9
1 1 1
so = 0.05555 · · · , and = 0.16666 · · · , so = 0.00166 · · · . Thus we can calculate:
18 6 600
1.00166
−0.05555
0.94611
50
51. Which of the following abelian groups of order 360 is NOT isomorphic to the other four?
(A) Z2 ⊕ Z2 ⊕ Z2 ⊕ Z3 ⊕ Z3 ⊕ Z5
(B) Z2 ⊕ Z2 ⊕ Z6 ⊕ Z15
(C) Z2 ⊕ Z3 ⊕ Z6 ⊕ Z10
(D) Z3 ⊕ Z4 ⊕ Z5 ⊕ Z6
(E) Z2 ⊕ Z6 ⊕ Z30
Solution:
The direct sum Zm ⊕ Zn is isomorphic to Zmn if and only if m and n are coprime. This means
that, in particular, Z2 ⊕ Z2 ∼
6 Z4 (it’s actually isomorphic to V4 instead). Thus choice D is the
=
odd one out. All of the other choices are produced by only combining coprime orders:
Z2 ⊕ Z2 ⊕ (Z2 ⊕ Z3 ) ⊕ (Z3 ⊕ Z5 ) ∼
= Z2 ⊕ Z2 ⊕ Z6 ⊕ Z15
Z2 ⊕ Z3 ⊕ (Z2 ⊕ Z3 ) ⊕ (Z2 ⊕ Z5 ) ∼
= Z2 ⊕ Z3 ⊕ Z6 ⊕ Z10
Z2 ⊕ (Z2 ⊕ Z3 ) ⊕ (Z2 ⊕ Z3 ⊕ Z5 ) ∼
= Z2 ⊕ Z6 ⊕ Z30
51
52. Let f : (0, 1) → (0, ∞) be a uniformly continuous function. Which of the following statements
are true?
I. If {xn } is a Cauchy sequence in (0, 1), then {f (xn )} is a Cauchy sequence in (0, ∞).
II. If lim xn exists, then lim f (xn ) = f lim xn .
n→∞ n→∞ n→∞
III. If {xn } and {yn } are two Cauchy sequences in (0, 1), then f (xn ) − f (yn ) is a Cauchy
sequence in (0, ∞).
(A) I only (B) I and II only (C) I and III only (D) II and III only (E) I, II, and III
Solution:
I. This is a basic fact from real analysis: uniform continuity preserves Cauchy sequences.
Let ε > 0; since f is uniformly continuous, there exists a δ > 0 such that |x − y| < δ
implies |f (x) − f (y)| < ε. Then, since {xn } is Cauchy, there exists an N > 0 such that
m, n > N implies |xm − xn | < δ, which then implies |f (xm ) − f (xn )| < ε. Thus {f (xn )}
is Cauchy.
II. Let lim xn = x, and let ε > 0. Since f is continuous, there exists a δ > 0 such that
n→∞
|xn − x| < δ implies |f (xn ) − f (x)| < ε. By the definition of limit, for this δ, there exists
an N > 0 such that n > N implies |xn − x| < δ, which then implies |f (xn ) − f (x)| < ε.
Thus lim f (xn ) = f (x). (This is actually just a property of continuous functions, as
n→∞
long as lim f (xn ) exists in the first place, which is what uniform continuity guarantees
n→∞
here.)
III. Let ε > 0. By I above, we know that {f (xn )} and {f (yn )} are also Cauchy sequences.
ε
This means that we can find an N1 such that m, n > N1 implies |f (xm ) − f (xn )| <
2
ε
and an N2 such that m, n > N2 implies |f (ym ) − f (yn )| < ; letting N = max {N1 , N2 },
2
both implications are satisfied. Now we can use the Triangle Inequality:
f (xm ) − f (ym ) − f (xn ) − f (yn ) = f (xm ) − f (ym ) − f (xn ) + f (yn )
= f (xm ) − f (xn ) − f (ym ) − f (yn )
ε ε
< + =ε
2 2
Thus f (xn ) − f (yn ) is a Cauchy sequence as well.
52
I
1
53. Calculate tan z dz, where C is the circle |z − 1| = 1 parametrized counterclockwise.
2πi C
π
This means that we just need to calculate the residue of tan z at z0 = . Doing so from the
2
p(z)
definition would be annoying, but we can use a shortcut formula: If f (z) = , where
q(z)
p(z0 )
p(z0 ) 6= 0, q(z0 ) = 0, and q 0 (z0 ) 6= 0, then Res f (z) = 0 . In this case we have p(z) = sin z
z=z0 q (z0 )
π
sin 2
and q(z) = cos z, so Resπ tan z = = −1.
z= 2 − sin π2
I I
1
From here, we know that tan z dz = 2πi · Resπ tan z, so tan z dz = Resπ tan z = −1.
C z= 2 2πi C z= 2
53
54. Suppose A is a 3 × 3 matrix with the property that A3 = A. Which of the following must be
true?
54
x 0 1 2 3 4
f (x) 1 −5 −7 −2 3
55. Selected values of a polynomial f (x) of degree 4 are given in the table above. What is the
value of f (5)?
x 0 1 2 3 4
f (x) 1 −5 −7 −2 3
∆f (x) −6 −2 5 5
2
∆ f (x) 4 7 0
∆3 f (x) 3 −7
∆4 f (x) −10
From the table we see that ∆4 f (x) = −10, so we build the value of f (5) backwards from here:
x 0 1 2 3 4 5
f (x) 1 −5 −7 −2 3 −9
∆f (x) −6 −2 5 5 −12
∆2 f (x) 4 7 0 −17
3
∆ f (x) 3 −7 −17
∆4 f (x) −10 −10
55
56. Let g(z) be an analytic function such that g(z) = 3 for all z in the open disk |z| < 2. If
g(1) = 3i, find all possible values of g(−1).
(A) {3i}
(B) {−3i}
(C) {3i, −3i}
(D) {3, −3, 3i, −3i}
(E) {z ∈ C : |z| = 3}
Solution:
In complex analysis, there are plenty of statements that boil down to “if an analytic function
has [some property], then it must be constant.” One such property is having a constant
modulus in an open disk. There are many ways to show this; this is one of them.
Suppose g(x+iy) = u(x, y)+iv(x, y), where u and v are real-valued functions of real variables
x and y. Then |g(x, y)|2 = u(x, y)2 + v(x, y)2 , which we are claiming is a constant. This means
that their partial derivatives must be zero:
∂ 2 ∂ 2
(u + v 2 ) = 0 (u + v 2 ) = 0
∂x ∂y
∂u ∂v ∂u ∂v
2u + 2v =0 2u + 2v =0
∂x ∂x ∂y ∂y
∂u ∂v ∂u ∂v
u +v =0 u +v =0
∂x ∂x ∂y ∂y
∂u ∂v
Now, the Cauchy-Riemann equations tell us that for an analytic function, = and
∂x ∂y
∂u ∂v
= − . We can use these to rewrite our equations above entirely in terms of v:
∂y ∂x
∂v ∂v ∂v ∂v
u +v =0 −u +v =0
∂y ∂x ∂x ∂y
Multiplying the first equation by u and the second equation by v and then adding them
together, we end up with:
∂v
(u2 + v 2 ) =0
∂y
∂v
If u2 + v 2 = 0, then g(z) = 0, so g is constant. Otherwise, if = 0, which means that v is
∂y
∂v
constant and therefore = 0 as well. Applying the Cauchy-Riemann equations once more,
∂x
we can show that u is constant as well, and that therefore g must be constant.
Thus we must have g(−1) = 3i.
56
θ
57. Let C1 be the curve defined by the polar equation r = for θ ≥ 0, and let C2 be the circle
θ+1
defined by the equation x2 + y 2 = 1. Let C = C1 ∪ C2 .
Which of the following statements are true with respect to the standard topology on R2 ?
I. For any point P ∈ C, there exists an open disk centered at P containing some point
Q ∈ C such that P 6= Q.
S
II. For any collection F of open disks such that C ⊆ F, there exists a finite subcollection
G ⊆ F such that C ⊆ G.
III. For any pair of points P ∈ C and Q ∈ C, there exists a continuous function f : [0, 1] → C
such that f (0) = P and f (1) = Q.
(A) I only (B) II only (C) I and II only (D) II and III only (E) I, II, and III
Solution:
The graph of C1 is a spiral that approaches the unit circle C2 , as shown in the figure below.
The resulting hsape is sometimes called the “topologist’s whirlpool”, and it’s very similar to
the well-known “topologist’s sine curve.”
I. This is saying that every point of C is a limit point of C; this is true, since C is the closure
of C1 and is therefore closed.
II. This is saying that C is compact; since C is closed and bounded, the Heine-Borel theorem
tells us that C is compact with respect to the standard topology on R2 .
III. This is saying that C is path-connected; however, much like the topologists’s sine curve,
while C is connected, it is NOT path-connected. Actually proving this rigorously on
the exam isn’t necessary — you just need to see the similarities between the curves and
make a quick judgment. If you want to see a proof, though, there’s one provided in
Introduction to Topology, Applied and Pure by Adams and Franzosa.
57
Z ∞
1
58. dx =
0 1 + eax
1 1 log 2 a
(A) (B) a log 2 (C) (D) (E)
a a log 2 a log 2
Solution:
Letting u = eax , we have du = aeax dx. We don’t have an extra eax anywhere, but that’s fine;
du
we can make the same substitution again and write du = au dx, or = dx. Hence we have:
au
Z ∞ Z ∞
1 ∞
Z
1 1 du 1
ax
dx = · = du
0 1 + e 1 1 + u au a 1 u(u + 1)
1 1 1
The fraction can be shown via partial fractions to equal − , which we can
u(u + 1) u u+1
use to integrate:
1 ∞ 1
Z
1 1 ∞
− du = log u − log(u + 1)
a 1 u u+1 a 1
1 u ∞
= log
a u+1 1
1 1
= log 1 − log
a 2
1
= (0 + log 2)
a
log 2
=
a
A slicker way to do this would be to multiply the top and bottom through by e−ax .
58
59. To the nearest thousand, approximately how many roots does the function f (x) = e−x sin(x2 )
have on the interval [0, 100]?
(A) 1000 (B) 2000 (C) 3000 (D) 4000 (E) 5000
Solution:
√ √ √
The zeros of sin(x2 ) will occur at the square roots of the zeros of sin x: x = 0, π, 2π, 3π,
√
and so on. Therefore we need to estimate the value of n such that nπ ≈ 100.
√
nπ ≈ 100
nπ ≈ 10 000
10 000
n≈
π
10 000
/
3
≈ 3 333
59
60. Circle C is tangent to the graph of y = x2 at the origin and has the same curvature as the
parabola at the point of tangency. What is the radius of circle C?
1 1 2 3
(A) (B) (C) (D) (E) 2
3 2 3 2
Solution:
Every GRE has a problem or two that uses some random formula you’ve likely forgotten. In
this case, it’s the curvature formula.
One common formula for curvature (that doesn’t require that we reparametrize in terms of
arclength) is:
kT0 (t)k
κ=
kr0 (t)k
√
Parametrizing our parabola as r(t) = (t, t2), we have r0 (t) = (1,2t), so kr0 (t)k = 1 + 4t2 .
r0 (t) 1 2t
Then we can calculate T(t) = 0 (t)k
= √ ,√ ; differentiating again, we
kr 1 +4t 2 1 + 4t2
4t 2
end up with T0 (t) = − 3/2
, .
2
(1 + 4t ) (1 + 4t2 )3/2
kT0 (0)k
Now r0 (0) = (1, 0) and T0 (0) = (0, 2), so κ = = 2. The osculating circle therefore
kr0 (0)k
1 1
has radius r = = .
κ 2
However, since we’re working in 2D, there’s an even faster formula we can use when y is a
function of x:
|y 00 |
κ=
(1 + (y 0 )2 )3/2
Since y 0 = 2x and y 00 = 2, this becomes:
|2|
κ= =2
(1 + (0)2 )3/2
1
Again we have κ = 2, so r = .
2
60
61. Let V be the vector space of continuous functions C → C under pointwise addition and
scalar multiplication by complex numbers, and let A be the set of fourth roots of unity in the
complex plane. For each α ∈ A, define the set Sα = {cos αz, sin αz, eαz }.
[
What is the dimension of the subspace of V spanned by Sα ?
α∈A
S1 = {cos z, sin z, ez }
Si = cos iz, sin iz, eiz = cosh z, i sinh z, eiz
From here, we notice immediately that by ignoring complex scalar multiples, our subspace
is spanned by eight functions:
61
62. Which of the following must be true of a function f : R2 → R?
∂f ∂f
I. If and exist and are continuous at (0, 0), then f is differentiable there.
∂x ∂y
II. If f has directional derivatives in all directions at (0, 0), then f is differentiable there.
∂2f ∂2f
III. If and exist at (0, 0), then they are equal there.
∂x∂y ∂y∂x
(A) None (B) II only (C) I and II only (D) I and III only (E) II and III only
Solution:
xy
if (x, y) 6= (0, 0)
I. Let f (x, y) = x2 + y2 . Then f is identically zero along both the
0 if (x, y) = (0, 0)
∂f ∂f
x-axis and the y-axis, so = = 0. However, the limit of f along the line y = x is
∂x ∂y
1
, so f is not even continuous, much less differentiable.
2
p
II. Let f (x, y) = 3 x2 y. Then if u = (cos θ, sin θ) is a unit vector, the directional derivative
f (h cos θ, h sin θ) − f (0, 0) √
3
of f at (0, 0) is fu0 (0, 0) = lim = cos2 θ sin θ, which exists for
h→0 h
∂f ∂f
any θ. We can get the partial derivatives and in particular by letting θ = 0 and
∂x ∂y
π
θ = , respectively, both of which give 0. However, this would imply that the directional
2
derivative equals ∇f (0, 0) · u = 0, but our expression in terms of θ is not always 0. Thus
f is not differentiable.
To visualize f , see https://www.math.tamu.edu/~tom.vogel/gallery/node17.html.
3 3
x y − xy if (x, y) 6= (0, 0)
III. Let f (x, y) = x2 + y 2 . Then, using the quotient rule, the partial
0 if (x, y) = (0, 0)
derivatives are:
∂f x4 y + 4x2 y 3 − y 5 ∂f x5 − 4x3 y 2 − xy 4
= =
∂x (x2 + y 2 )2 ∂y (x2 + y 2 )2
f (x, 0) − f (0, 0)
Notice that setting x = 0 gives us fx (0, y) = −y, so fx (0, 0) = lim = 0.
x→0 x
Similarly, fy (x, 0) = x and fy (0, 0) = 0. Differentiating with respect to the other variable,
∂ 2 f 2
0 (0, 0) = lim fx (0, y) − fx (0, 0) = −1, but ∂ f
we have = fxy =
∂y∂x (x,y)=(0,0) y ∂x∂y (x,y)=(0,0)
y→0
0 (0, 0) = lim fx (x, 0) − fx (0, 0) = 1. Thus the mixed partials are not equal despite
fyx
x→0 x
both being defined.
62
1 2
− 23
3 3
63. Consider the matrix equation QRx = b, where Q = − 23 2 1
is an orthogonal matrix,
3 3
2 1 2
3 3 3
3 1 4 11
R = 0 1 5 is an upper triangular matrix, and b = −10.
0 0 9 −65
x1
If x = x2 , what is the value of x3 ?
x3
We can get the value of x3 by multiplying the last row of each matrix by the column vector
on each side of the equation:
63
64. Let A and B be ideals of a ring R, and define the ideals A + B and AB as follows:
A + B = a + b : a ∈ A, b ∈ B
AB = a1 b1 + · · · + an bn : ai ∈ A, bi ∈ B, i ∈ {1, . . . , n} , n ∈ {1, 2, . . .}
64
65. For each pair of integers a and b, let the set {a + bn : n ∈ Z} be denoted a + bZ. Consider
the topology T on Z whose basis is {a + bZ : a, b ∈ Z and b 6= 0}. Which of the following
statements is false?
The union of these open sets is Z \ {0}. Since the union of any number of open sets in a
topology is open, Z \ {0} is open, so {0} must be closed.
D: Since the basis elements are all infinite, the only way we could possibly get a finite
open set is by an intersection. It suffices to consider what happens when intersecting
basis elements: if two basis elements a + bZ and c + dZ have at least one element k in
common, then they will again overlap after any multiple of lcm(b, d), so we can explicitly
calculate that (a + bZ) ∩ (c + dZ) = k + lcm(b, d)Z. Since we’re only allowed to take finite
intersections, all open sets must be infinite.
E: This may be true on the standard topology on R, but it’s not true here! For example, the
set 1 + 2Z is open since it’s a basis element, but since its complement 0 + 2Z is also open,
1 + 2Z is closed as well. In fact, it can be shown that any basis element is both open and
closed!
65
66. Let an abelian group M form a left module over a ring R. We say that a subset S of M is a
spanning set of M if for every m ∈ M , there exist {r1 , r2 , . . . , rn } ⊆ R and {s1 , s2 , . . . , sn } ⊆ S
such that
Xn
m= ri si .
i=1
We call this spanning set S minimal if no proper subset of S is a spanning set for S. In
addition, S is called a basis for M if m = 0 implies r1 = r2 = · · · = rn = 0.
Which of the following statements must be true?
I. If R is finite, then any basis S of M is finite.
II. If S and S 0 are two distinct minimal spanning sets of M , then |S| = |S 0 |.
III. If R is a field and S is a finite minimal spanning set of M , then S is a basis of M .
(A) II only (B) III only (C) I and II only (D) I and III only (E) II and III only
Solution:
You can think of a module intuitively as being just like a vector space, with the elements of
the group M being like the “vectors,” except instead of the “scalar” coefficients coming from
a field, they come from a ring.
I. One example of a module is R[x], the set of all polynomials with coefficients in a ring R.
Even if R is finite — say, perhaps R = Z5 — the basis 1, x, x2 , x3 , · · · is infinite.
II. Another example of a module is any abelian group, which can be turned into a module
using Z as the ring of coefficients. In fact, we can even make the group be Z as well, so
we’re just multiplying and adding integers.
Now, an easy spanning set for Z would be S = {1}, since every integer can be written as
a · 1 for some a ∈ Z. This spanning set is minimal, since if you take 1 out, there’s nothing
left!
A different spanning set would be S 0 = {2, 3}, since gcd(2, 3) = 1 and therefore every
integer can be written as a · 2 + b · 3 for some a, b ∈ Z. This spanning set is also minimal,
since removing 2 would only span 3Z and removing 3 would only span 2Z.
6 |S 0 |. This is a way in which modules are different from vector spaces —
However, |S| =
they don’t necessarily have the invariant basis number condition.
III. If R is a field, then M is actually a vector space after all! Vector spaces do have the
invariant basis number condition. Since M has a finite minimal spanning set S, that set
S must be a basis. (It’s a good thing we have the finiteness condition given, because
otherwise we’d need to invoke Zorn’s Lemma to claim that every vector space has a
basis!)
Many thanks to the folks at https://discord.sg/math for helping me put together this question!
66