IIT JAM 2025 Mathematical Statistics Question Paper with Answer Key PDF Available

Shivam Yadav's profile photo

Shivam Yadav

Educational Content Expert | Updated on - Sep 9, 2025

IIT JAM 2025 Mathematical Statistics (MS) held today, February 2, 2025, in Shift 2, between 2:30 PM and 5:30 PM has now been concluded.

The Exam was of moderate to high difficulty level, featuring both theoretical and application-based problems and it consisted of 60 questions, divided into 30 MCQs, 10 MSQs, and 20 NATs, carrying a total of 100 marks, with negative marking applicable only for MCQs.

The IIT JAM 2025 Mathematical Statistics Question Paper and the Answer Key will be available for download soon.

Also, Check: IIT JAM 2025 Question Paper

IIT JAM 2025 Mathematical Statistics (MS) Question Paper with Answer Key PDFs

IIT JAM 2025 Mathematical Statistics (MS) Question Paper with Answer Key PDFs Download PDF Check Solutions

IIT JAM 2025 Mathematical Statistics (MS) Question Paper with Solutions


Question 1:

Let \(\{a_n\}_{n \ge 1}\) and \(\{b_n\}_{n \ge 1}\) be sequences given by \[ a_n = \left\lfloor \frac{n^2}{n+1} \right\rfloor \quad and \quad b_n = \frac{n^2}{n+1} - a_n \,. \]
Then

  • (A) \(\{a_n\}_{n \ge 1}\) converges and \(\{b_n\}_{n \ge 1}\) diverges
  • (B) \(\{a_n\}_{n \ge 1}\) diverges and \(\{b_n\}_{n \ge 1}\) converges
  • (C) Both \(\{a_n\}_{n \ge 1}\) and \(\{b_n\}_{n \ge 1}\) diverge
  • (D) Both \(\{a_n\}_{n \ge 1}\) and \(\{b_n\}_{n \ge 1}\) converge
Correct Answer: (B) \(\{a_n\}_{n \ge 1}\) diverges and \(\{b_n\}_{n \ge 1}\) converges
View Solution




Step 1: Understanding the Concept:

To determine the convergence or divergence of the sequences, we need to find their explicit formulas and then evaluate their limits as \( n \to \infty \). A sequence converges if its limit is a finite real number; otherwise, it diverges.




Step 2: Analyzing the sequence \(\{a_n\}\):

First, we simplify the expression inside the floor function for \(a_n\). We use polynomial division or algebraic manipulation:
\[ \frac{n^2}{n+1} = \frac{n^2 - 1 + 1}{n+1} = \frac{(n-1)(n+1) + 1}{n+1} = (n-1) + \frac{1}{n+1} \]
Now, we can write \(a_n\) as:
\[ a_n = \left\lfloor (n-1) + \frac{1}{n+1} \right\rfloor \]
For \( n \ge 1 \), we know that \(n-1\) is an integer. Also, \( 0 < \frac{1}{n+1} \le \frac{1}{2} \).

The floor of an integer plus a small positive fraction (less than 1) is the integer itself.

Therefore, \( a_n = n-1 \).

To check for convergence, we take the limit as \( n \to \infty \):
\[ \lim_{n \to \infty} a_n = \lim_{n \to \infty} (n-1) = \infty \]
Since the limit is not finite, the sequence \(\{a_n\}\) diverges.




Step 3: Analyzing the sequence \(\{b_n\}\):

The sequence \(\{b_n\}\) is defined as \(b_n = \frac{n^2}{n+1} - a_n\).

Using the results from Step 2, we substitute the expressions for \(\frac{n^2}{n+1}\) and \(a_n\):
\[ b_n = \left( (n-1) + \frac{1}{n+1} \right) - (n-1) \] \[ b_n = \frac{1}{n+1} \]
To check for convergence, we take the limit as \( n \to \infty \):
\[ \lim_{n \to \infty} b_n = \lim_{n \to \infty} \frac{1}{n+1} = 0 \]
Since the limit is a finite number (0), the sequence \(\{b_n\}\) converges.




Step 4: Final Answer:

The sequence \(\{a_n\}\) diverges, and the sequence \(\{b_n\}\) converges. This corresponds to option (B).
Quick Tip: When dealing with a floor function of a rational expression, polynomial long division is a powerful tool to separate the integer part from the fractional part. The fractional part of a number \(x\) is given by \(x - \lfloor x \rfloor\). Here, \(b_n\) is precisely the fractional part of \(\frac{n^2}{n+1}\).


Question 2:

Let \(a, b, c\) be real numbers with \(b \ne c\). Define the matrix

Then the number of characteristic roots of \(M\) that are real is

  • (A) 3
  • (B) 2
  • (C) 1
  • (D) 0
Correct Answer: (C) 1
View Solution




Step 1: Understanding the Concept:

The characteristic roots (or eigenvalues) of a matrix \(M\) are the solutions \(\lambda\) to the characteristic equation \(\det(M - \lambda I) = 0\). For a real matrix like \(M\), the characteristic polynomial is a polynomial with real coefficients. Its roots can be real or complex. Complex roots of a real polynomial must occur in conjugate pairs. Since \(M\) is a \(3 \times 3\) matrix, its characteristic polynomial is cubic, which must have at least one real root. Thus, the number of real roots can be 1 or 3.




Step 2: Key Formula or Approach:

The given matrix \(M\) is a circulant matrix. The eigenvalues of a \(3 \times 3\) circulant matrix of the form


are given by the formula \(\lambda_k = c_0 + c_1 \omega^k + c_2 \omega^{2k}\) for \(k = 0, 1, 2\), where \(\omega = e^{i2\pi/3}\) is a primitive cube root of unity.

The cube roots of unity are \(1, \omega = -\frac{1}{2} + i\frac{\sqrt{3}}{2}, and \omega^2 = -\frac{1}{2} - i\frac{\sqrt{3}}{2}\).




Step 3: Detailed Explanation:

For our matrix \(M\), we have \(c_0 = a, c_1 = b, c_2 = c\). Let's find the three eigenvalues:


For \(k=0\):

\(\lambda_0 = a + b\omega^0 + c\omega^0 = a + b(1) + c(1) = a+b+c\).

Since \(a, b, c\) are real, \(\lambda_0\) is a real eigenvalue.

For \(k=1\):

\(\lambda_1 = a + b\omega^1 + c\omega^2 = a + b\left(-\frac{1}{2} + i\frac{\sqrt{3}}{2}\right) + c\left(-\frac{1}{2} - i\frac{\sqrt{3}}{2}\right)\).

\(\lambda_1 = \left(a - \frac{b}{2} - \frac{c}{2}\right) + i\left(\frac{b\sqrt{3}}{2} - \frac{c\sqrt{3}}{2}\right) = \left(a - \frac{b+c}{2}\right) + i\frac{\sqrt{3}}{2}(b-c)\).

For \(k=2\):

\(\lambda_2 = a + b\omega^2 + c\omega^4 = a + b\omega^2 + c\omega\). (Since \(\omega^3 = 1\))

\(\lambda_2 = a + b\left(-\frac{1}{2} - i\frac{\sqrt{3}}{2}\right) + c\left(-\frac{1}{2} + i\frac{\sqrt{3}}{2}\right)\).

\(\lambda_2 = \left(a - \frac{b+c}{2}\right) - i\frac{\sqrt{3}}{2}(b-c)\).

The eigenvalues \(\lambda_1\) and \(\lambda_2\) are complex conjugates. They are real if and only if their imaginary part is zero. The imaginary part of \(\lambda_1\) is \(\frac{\sqrt{3}}{2}(b-c)\).

This imaginary part is zero only if \(b-c=0\), i.e., \(b=c\).

However, the problem states that \(b \ne c\). Therefore, \(b-c \ne 0\), which means the imaginary part is non-zero.

Thus, \(\lambda_1\) and \(\lambda_2\) are non-real complex eigenvalues.




Step 4: Final Answer:

The matrix \(M\) has one real eigenvalue (\(\lambda_0\)) and two non-real complex conjugate eigenvalues (\(\lambda_1, \lambda_2\)). Therefore, the number of real characteristic roots is exactly 1.
Quick Tip: An alternative method is to check if \(v = (1, 1, 1)^T\) is an eigenvector.

This confirms that \(a+b+c\) is a real eigenvalue. This quickly establishes that there is at least one real root.


Question 3:

Let \(f: \mathbb{R} \to \mathbb{R}\) be a continuous odd function that is not identically zero. Further, suppose that \(f\) is a periodic function. Define \[ g(x) = \int_0^x f(t) \, dt \,. \]
Then

  • (A) \(g\) is odd and not periodic
  • (B) \(g\) is odd and periodic
  • (C) \(g\) is even and not periodic
  • (D) \(g\) is even and periodic
Correct Answer: (D) \(g\) is even and periodic
View Solution




Step 1: Understanding the Concept:

We need to determine the properties (even/odd and periodic) of the function \(g(x)\), which is defined as the integral of another function \(f(x)\) with known properties.

An even function satisfies \(g(-x) = g(x)\).
An odd function satisfies \(g(-x) = -g(x)\).
A periodic function with period \(T\) satisfies \(g(x+T) = g(x)\) for some \(T>0\).




Step 2: Checking if \(g\) is Even or Odd:

We evaluate \(g(-x)\) using its definition: \[ g(-x) = \int_0^{-x} f(t) \, dt \]
Let's use the substitution \(u = -t\), which means \(t = -u\) and \(dt = -du\). The limits of integration change as follows: when \(t=0\), \(u=0\); when \(t=-x\), \(u=x\). \[ g(-x) = \int_{0}^{x} f(-u) \, (-du) = - \int_{0}^{x} f(-u) \, du \]
We are given that \(f\) is an odd function, so \(f(-u) = -f(u)\). \[ g(-x) = - \int_{0}^{x} (-f(u)) \, du = \int_{0}^{x} f(u) \, du \]
By definition, \(\int_{0}^{x} f(u) \, du = g(x)\).
Therefore, \(g(-x) = g(x)\), which means that \(g\) is an even function.




Step 3: Checking if \(g\) is Periodic:

Let \(T > 0\) be the period of \(f\), so \(f(t+T) = f(t)\) for all \(t\). We need to check if \(g(x+T) = g(x)\). \[ g(x+T) = \int_0^{x+T} f(t) \, dt = \int_0^x f(t) \, dt + \int_x^{x+T} f(t) \, dt \] \[ g(x+T) = g(x) + \int_x^{x+T} f(t) \, dt \]
For \(g\) to be periodic with period \(T\), the integral term must be zero for all \(x\).
The integral of a periodic function over any interval of length equal to its period is constant. So, \[ \int_x^{x+T} f(t) \, dt = \int_0^{T} f(t) \, dt \]
Now we need to evaluate \(\int_0^{T} f(t) \, dt\). We use the property \(\int_0^a h(t) \, dt = \int_0^a h(a-t) \, dt\).
Let \(I = \int_0^T f(t) \, dt\). Then, \[ I = \int_0^T f(T-t) \, dt \]
Since \(f\) is periodic with period \(T\), \(f(T-t) = f(-t)\). \[ I = \int_0^T f(-t) \, dt \]
Since \(f\) is an odd function, \(f(-t) = -f(t)\). \[ I = \int_0^T -f(t) \, dt = - \int_0^T f(t) \, dt = -I \]
So, we have \(I = -I\), which implies \(2I=0\), and thus \(I=0\).
Since \(\int_0^T f(t) \, dt = 0\), we have \(g(x+T) = g(x) + 0 = g(x)\).
This means that \(g\) is a periodic function.




Step 4: Final Answer:

We have shown that \(g(x)\) is both an even function and a periodic function. This corresponds to option (D).
Quick Tip: Remember these general rules: The definite integral of an odd function over a symmetric interval \([-a, a]\) is always zero. The integral of an odd periodic function over one full period is always zero. The integral of an odd function from 0 to \(x\) results in an even function. The integral of an even function from 0 to \(x\) results in an odd function.


Question 4:

Suppose \(Z_1, Z_2, \dots, Z_{128}\) are i.i.d. Bin(1, 0.5) random variables. Define \[ \mathbf{X} = (Z_1, Z_2, \dots, Z_{64})^T \quad and \quad \mathbf{Y} = (Z_{65}, Z_{66}, \dots, Z_{128})^T \,. \]
Then the value of Var(\(\mathbf{X}^T\mathbf{Y}\)) is

  • (A) 4
  • (B) 8
  • (C) 12
  • (D) 16
Correct Answer: (C) 12
View Solution






Step 1: Understanding the Concept:

We need to find the variance of the scalar product \(S = \mathbf{X}^T\mathbf{Y}\). The variables \(Z_i\) are i.i.d. Bernoulli variables with \(p=0.5\), since Bin(1, p) is the Bernoulli(p) distribution.
For each \(Z_i\):

Expected value: \(E[Z_i] = p = 0.5\)
Variance: \(Var(Z_i) = p(1-p) = 0.5(0.5) = 0.25\)

The scalar product is given by: \[ S = \mathbf{X}^T\mathbf{Y} = \sum_{i=1}^{64} Z_i Z_{64+i} \]



Step 2: Key Formula or Approach:

Let \(W_i = Z_i Z_{64+i}\). Then \(S = \sum_{i=1}^{64} W_i\).
The indices used for each \(W_i\) (i.e., \(\{i, 64+i\}\)) are disjoint from the indices for any other \(W_j\) (where \(j \ne i\)). Since all \(Z_k\) are i.i.d., the random variables \(W_i\) are also i.i.d.
Therefore, the variance of the sum is the sum of the variances: \[ Var(S) = Var\left(\sum_{i=1}^{64} W_i\right) = \sum_{i=1}^{64} Var(W_i) = 64 \cdot Var(W_1) \]
To find Var(\(S\)), we first need to find the variance of a single term \(W_1 = Z_1 Z_{65}\).



Step 3: Detailed Explanation:

First, let's analyze the variable \(W_i = Z_i Z_{64+i}\). Since \(Z_i\) and \(Z_{64+i}\) can only take values 0 or 1, their product \(W_i\) can also only be 0 or 1. This means \(W_i\) is also a Bernoulli random variable. Let's find its parameter \(p_W\). \[ p_W = P(W_i = 1) = P(Z_i = 1 and Z_{64+i} = 1) \]
Because \(Z_i\) and \(Z_{64+i}\) are independent: \[ p_W = P(Z_i = 1) \cdot P(Z_{64+i} = 1) = (0.5) \cdot (0.5) = 0.25 \]
So, each \(W_i\) is a Bernoulli(0.25) random variable. The variance of \(W_i\) is: \[ Var(W_i) = p_W(1-p_W) = 0.25(1 - 0.25) = 0.25 \times 0.75 = \frac{1}{4} \times \frac{3}{4} = \frac{3}{16} \]
Now we can calculate the variance of \(S\): \[ Var(S) = 64 \cdot Var(W_1) = 64 \times \frac{3}{16} = 4 \times 3 = 12 \]



Step 4: Alternative Method (Law of Total Variance):

Var(\(S\)) = E[Var(\(S|\mathbf{X}\))] + Var(E[\(S|\mathbf{X}\)]).

\(E[S|\mathbf{X}] = E[\sum Z_i Z_{64+i} | \mathbf{X}] = \sum Z_i E[Z_{64+i}] = \sum Z_i (0.5) = 0.5 \sum Z_i\). Let \(K = \sum_{i=1}^{64} Z_i \sim Bin(64, 0.5)\). So \(E[S|\mathbf{X}] = 0.5K\).
Var(E[\(S|\mathbf{X}\)]) = Var(\(0.5K\)) = \(0.25 \cdot Var(K) = 0.25 \cdot (64 \times 0.5 \times 0.5) = 0.25 \cdot 16 = 4\).
Var(\(S|\mathbf{X}\)) = Var(\(\sum Z_i Z_{64+i} | \mathbf{X}\)) = \(\sum Var(Z_i Z_{64+i} | \mathbf{X})\) = \(\sum Z_i^2 Var(Z_{64+i})\). Since \(Z_i \in \{0,1\}\), \(Z_i^2 = Z_i\).
Var(\(S|\mathbf{X}\)) = \(\sum Z_i \cdot Var(Z_{64+i}) = \sum Z_i \cdot (0.25) = 0.25K\).
E[Var(\(S|\mathbf{X}\))] = E[\(0.25K\)] = \(0.25 \cdot E[K] = 0.25 \cdot (64 \times 0.5) = 0.25 \cdot 32 = 8\).

Var(\(S\)) = \(8 + 4 = 12\).

Both methods confirm the answer is 12. Quick Tip: If the question were to ask for Var(\(\mathbf{X}^T\mathbf{X}\)), the calculation would be different. In that case, \(S' = \mathbf{X}^T\mathbf{X} = \sum_{i=1}^{64} Z_i^2\). Since \(Z_i\) is a Bernoulli variable, \(Z_i^2 = Z_i\). Thus, \(S' = \sum_{i=1}^{64} Z_i\), which follows a Binomial(64, 0.5) distribution. The variance would then be \(np(1-p) = 64 \times 0.5 \times 0.5 = 16\). Given the options, it's possible this was the intended question. However, based on the literal text, the answer is 12.


Question 5:

Let \(X_1, X_2, X_3\) be i.i.d. Bin(1, \(\theta\)) random variables. Consider the problem of testing the null hypothesis \(H_0: \theta = \frac{1}{2}\) against the alternative hypothesis \(H_1: \theta = \frac{1}{4}\) based on \(X_1, X_2, X_3\). Then the power of the most powerful test of size 0.125 is

  • (A) 0
  • (B) \( \frac{1}{64} \)
  • (C) \( \frac{27}{64} \)
  • (D) \( \frac{7}{8} \)
Correct Answer: (C) \( \frac{27}{64} \)
View Solution




Step 1: Understanding the Concept:

We need to find the power of the Most Powerful (MP) test for a simple null hypothesis against a simple alternative. The Neyman-Pearson Lemma provides the form of the MP test. The test statistic is the likelihood ratio.
Let \(L(\theta; \mathbf{x})\) be the likelihood function. The MP test rejects \(H_0\) if the likelihood ratio \(\frac{L(\theta_1; \mathbf{x})}{L(\theta_0; \mathbf{x})} > k\) for some constant \(k\). Here, \(\theta_0 = 1/2\) and \(\theta_1 = 1/4\).



Step 2: Constructing the Most Powerful Test:

The random variables are \(X_i \sim Bernoulli(\theta)\). The likelihood function is: \[ L(\theta; x_1, x_2, x_3) = \prod_{i=1}^3 \theta^{x_i} (1-\theta)^{1-x_i} = \theta^{\sum x_i} (1-\theta)^{3 - \sum x_i} \]
Let \(S = \sum_{i=1}^3 X_i\). Under both \(H_0\) and \(H_1\), \(S\) is a sufficient statistic for \(\theta\), and \(S \sim Bin(3, \theta)\). The test can be based on \(S\).
The likelihood ratio is: \[ \frac{L(\theta_1=\frac{1}{4}; S)}{L(\theta_0=\frac{1}{2}; S)} = \frac{(\frac{1}{4})^S (1-\frac{1}{4})^{3-S}}{(\frac{1}{2})^S (1-\frac{1}{2})^{3-S}} = \frac{(\frac{1}{4})^S (\frac{3}{4})^{3-S}}{(\frac{1}{2})^S (\frac{1}{2})^{3-S}} = \frac{3^{3-S}}{4^3} \cdot \frac{2^3}{1} = \frac{3^{3-S}}{8} \]
The MP test rejects \(H_0\) if \(\frac{3^{3-S}}{8} > k\), or equivalently \(3^{3-S} > 8k\).
Since \(3^{3-S}\) is a decreasing function of \(S\), this is equivalent to rejecting \(H_0\) if \(S < c\) for some constant \(c\).
The critical region of the MP test is of the form \(C = \{\mathbf{x} : S < c\}\).



Step 3: Determining the Critical Region using the Size:

The size of the test is the probability of Type I error, \(\alpha = P(Reject H_0 | H_0 is true)\).
We are given \(\alpha = 0.125 = 1/8\).
Under \(H_0\), \(\theta = 1/2\), so \(S \sim Bin(3, 1/2)\). The possible values for \(S\) are 0, 1, 2, 3.
The probability mass function under \(H_0\) is \(P(S=s) = \binom{3}{s} (\frac{1}{2})^s (\frac{1}{2})^{3-s} = \binom{3}{s} \frac{1}{8}\).

\(P(S=0|H_0) = \binom{3}{0}\frac{1}{8} = \frac{1}{8}\)
\(P(S=1|H_0) = \binom{3}{1}\frac{1}{8} = \frac{3}{8}\)
\(P(S=2|H_0) = \binom{3}{2}\frac{1}{8} = \frac{3}{8}\)
\(P(S=3|H_0) = \binom{3}{3}\frac{1}{8} = \frac{1}{8}\)

The critical region is of the form \(S < c\). We need to find \(c\) such that the size is 0.125.
Let's test possible values for \(c\):
If the critical region is \(S=0\), the size is \(P(S=0|H_0) = 1/8 = 0.125\).
This matches the given size. So, the critical region is \(C = \{\mathbf{x} : S=0\}\).
The MP test is to reject \(H_0\) if and only if \(S = X_1+X_2+X_3 = 0\).



Step 4: Calculating the Power of the Test:

The power of the test is the probability of correctly rejecting \(H_0\), which is \(P(Reject H_0 | H_1 is true)\).
Power = \(P(S \in C | H_1)\) = \(P(S=0 | \theta = 1/4)\).
Under \(H_1\), \(\theta = 1/4\), so \(S \sim Bin(3, 1/4)\).
The probability is: \[ P(S=0 | \theta = 1/4) = \binom{3}{0} \left(\frac{1}{4}\right)^0 \left(1-\frac{1}{4}\right)^{3-0} = 1 \cdot 1 \cdot \left(\frac{3}{4}\right)^3 = \frac{27}{64} \]
So, the power of the test is \(27/64\). Quick Tip: The Neyman-Pearson Lemma is fundamental for finding the Most Powerful test between two simple hypotheses. The key is to form the likelihood ratio, which often simplifies to a condition on a sufficient statistic. The size of the test determines the exact critical region, and the power is the probability of this region under the alternative hypothesis.


Question 6:

Suppose \(X\) is a Poisson(\(\lambda\)) random variable. Define \(Y = (-1)^X\). Then the expected value of \(Y\) is

  • (A) \(-\lambda e^{-2\lambda}\)
  • (B) \( -e^{-2\lambda} \)
  • (C) \(\lambda e^{-2\lambda}\)
  • (D) \({\lambda}\)
Correct Answer: (A) \(-\lambda e^{-2\lambda}\)
View Solution




Step 1: Understanding the Concept:

We need to find the expected value of a function of a discrete random variable. For a discrete random variable \(X\) with probability mass function (PMF) \(P(X=k)\), the expected value of a function \(g(X)\) is given by \(E[g(X)] = \sum_k g(k) P(X=k)\).




Step 2: Key Formula or Approach:

Here, \(X \sim Poisson(\lambda)\), so its PMF is \(P(X=k) = \frac{e^{-\lambda}\lambda^k}{k!}\) for \(k = 0, 1, 2, \dots\).
The function is \(g(X) = Y = (-1)^X\).
The expected value is: \[ E[Y] = E[(-1)^X] = \sum_{k=0}^{\infty} (-1)^k P(X=k) = \sum_{k=0}^{\infty} (-1)^k \frac{e^{-\lambda}\lambda^k}{k!} \]



Step 3: Detailed Explanation:

We can factor out the constant \(e^{-\lambda}\) from the summation: \[ E[Y] = e^{-\lambda} \sum_{k=0}^{\infty} \frac{(-1)^k \lambda^k}{k!} = e^{-\lambda} \sum_{k=0}^{\infty} \frac{(-\lambda)^k}{k!} \]
The summation is the Taylor series expansion for \(e^z\) evaluated at \(z = -\lambda\): \[ \sum_{k=0}^{\infty} \frac{(-\lambda)^k}{k!} = e^{-\lambda} \]
Substituting this back into the expression for \(E[Y]\): \[ E[Y] = e^{-\lambda} \cdot (e^{-\lambda}) = e^{-2\lambda} \]
Thus, the expected value of \(Y\) is \(e^{-2\lambda}\).



Step 4: Final Answer:

The calculated expected value is \(e^{-2\lambda}\). Quick Tip: This is a standard problem related to the moment generating function (MGF) or probability generating function (PGF). The PGF of a Poisson(\(\lambda\)) is \(G_X(z) = E[z^X] = e^{\lambda(z-1)}\). The expected value we want is \(E[(-1)^X]\), which is simply the PGF evaluated at \(z=-1\). \(E[Y] = G_X(-1) = e^{\lambda(-1-1)} = e^{-2\lambda}\).


Question 7:

Let \(\{Y_n\}_{n \ge 1}\) be a sequence of i.i.d. Bin(1, p) random variables, where \(0 < p < 1\) is an unknown parameter. Let \(\hat{p}_n\) be the maximum likelihood estimator of \(p\) based on \(Y_1, Y_2, \dots, Y_n\). It is claimed that:

Which of the following statements is correct?

  • (A) (I) is correct and (II) is incorrect
  • (B) (I) is incorrect and (II) is correct
  • (C) Both (I) and (II) are correct
  • (D) Both (I) and (II) are incorrect
Correct Answer: (C) Both (I) and (II) are correct
View Solution




Step 1: Understanding the Concept:

This question deals with the asymptotic properties of the Maximum Likelihood Estimator (MLE) for the parameter of a Bernoulli distribution. We need to identify the correct forms of the Central Limit Theorem (CLT) and its practical application using Slutsky's Theorem.




Step 2: Finding the MLE and Analyzing Statement (I):

The likelihood function for \(n\) Bernoulli trials is \(L(p) = p^{\sum y_i} (1-p)^{n - \sum y_i}\). The MLE for \(p\) is the sample mean: \[ \hat{p}_n = \frac{1}{n}\sum_{i=1}^n Y_i = \bar{Y}_n \]
The \(Y_i\) are i.i.d. with mean \(E[Y_i] = p\) and variance \(Var(Y_i) = p(1-p)\).
By the classical Central Limit Theorem (Lindeberg-Lévy CLT), the sample mean \(\bar{Y}_n\) is asymptotically normally distributed. Specifically, \[ \frac{\bar{Y}_n - E[\bar{Y}_n]}{\sqrt{Var(\bar{Y}_n)}} \xrightarrow{d} N(0,1) \]
Here, \(E[\bar{Y}_n] = p\) and \(Var(\bar{Y}_n) = \frac{Var(Y_i)}{n} = \frac{p(1-p)}{n}\).
Substituting \(\hat{p}_n\) for \(\bar{Y}_n\), we get: \[ \frac{\hat{p}_n - p}{\sqrt{\frac{p(1-p)}{n}}} \xrightarrow{d} N(0,1) \]
This matches statement (I). Therefore, (I) is correct.




Step 3: Analyzing Statement (II):

Statement (II) replaces the true variance \(p(1-p)\) in the denominator with its consistent estimate \(\hat{p}_n(1-\hat{p}_n)\). We need to determine if this replacement preserves the convergence in distribution.
By the Weak Law of Large Numbers (WLLN), the sample mean converges in probability to the true mean. \[ \hat{p}_n = \bar{Y}_n \xrightarrow{p} p \]
Since the function \(g(x) = x(1-x)\) is a continuous function, by the Continuous Mapping Theorem, \[ \hat{p}_n(1-\hat{p}_n) \xrightarrow{p} p(1-p) \]
Now we apply Slutsky's Theorem. We have two convergences:

\(X_n = \sqrt{n}(\hat{p}_n - p) \xrightarrow{d} N(0, p(1-p))\), which is equivalent to \(\frac{\hat{p}_n - p}{\sqrt{p(1-p)/n}} \xrightarrow{d} N(0,1)\)
\(Y_n = \sqrt{\frac{p(1-p)}{\hat{p}_n(1-\hat{p}_n)}} \xrightarrow{p} 1\)

Slutsky's Theorem states that if \(X_n \xrightarrow{d} X\) and \(Y_n \xrightarrow{p} c\), then \(X_n Y_n \xrightarrow{d} cX\).
Here, \(c=1\), so the product of our two sequences converges in distribution to \(N(0,1) \times 1 = N(0,1)\). \[ \left( \frac{\hat{p}_n - p}{\sqrt{\frac{p(1-p)}{n}}} \right) \cdot \left( \sqrt{\frac{p(1-p)}{\hat{p}_n(1-\hat{p}_n)}} \right) = \frac{\hat{p}_n - p}{\sqrt{\frac{\hat{p}_n(1-\hat{p}_n)}{n}}} \xrightarrow{d} N(0,1) \]
This matches statement (II). Therefore, (II) is correct.




Step 4: Final Answer:

Both statements (I) and (II) are correct statements about the asymptotic distribution of the MLE for a Bernoulli parameter. (I) is a direct result of the CLT, and (II) follows from the CLT combined with Slutsky's Theorem.
Quick Tip: Statement (I) is the direct result of the Central Limit Theorem. Statement (II) is the "practical" version of the CLT, used for constructing confidence intervals and hypothesis tests when the true parameter \(p\) is unknown and must be estimated from the data. The justification for this replacement is Slutsky's Theorem.


Question 8:

Let \(X\) be a continuous random variable with probability density function \(f(x)\). Consider the problem of testing the null hypothesis

against the alternative hypothesis

Then the power of the most powerful size \(\alpha\) test, where \(0 < \alpha < 1\), based on a single sample, is

  • (A) \(\alpha(1-\alpha)\)
  • (B) \(\alpha(2-\alpha)\)
  • (C) \(1-\alpha\)
  • (D) \(\alpha\)
Correct Answer: (B) \(\alpha(2-\alpha)\)
View Solution




Step 1: Understanding the Concept:

We need to find the power of the Most Powerful (MP) test of size \(\alpha\). The Neyman-Pearson Lemma provides the form of the MP test. The test is based on a single observation \(X\). The test rejects \(H_0\) if the likelihood ratio \(\frac{f_1(x)}{f_0(x)} > k\) for some constant \(k\).




Step 2: Constructing the Most Powerful Test:

The likelihood ratio based on a single observation \(x\) (where \(0 < x < 1\)) is: \[ \Lambda(x) = \frac{f_1(x)}{f_0(x)} = \frac{2x}{1} = 2x \]
The MP test rejects \(H_0\) if \(\Lambda(x) = 2x > k\), which is equivalent to rejecting \(H_0\) if \(x > k/2\). Let \(c = k/2\). The critical region (rejection region) is of the form \(C = \{x : x > c\}\) for some constant \(c \in (0,1)\).




Step 3: Determining the Critical Region using the Size \(\alpha\):

The size of the test is the probability of Type I error: \(\alpha = P(Reject H_0 | H_0 is true)\).
Under \(H_0\), \(X \sim U(0,1)\) (Uniform distribution on (0,1)). \[ \alpha = P(X \in C | H_0) = P(X > c | H_0) = \int_c^1 f_0(x) \, dx = \int_c^1 1 \, dx = [x]_c^1 = 1 - c \]
So, we have \(\alpha = 1 - c\), which implies \(c = 1 - \alpha\).
The critical region for the MP test of size \(\alpha\) is \(C = \{x : x > 1 - \alpha\}\).




Step 4: Calculating the Power of the Test:

The power of the test is the probability of correctly rejecting \(H_0\): Power = \(P(Reject H_0 | H_1 is true)\).
Under \(H_1\), the pdf of \(X\) is \(f_1(x) = 2x\) for \(0 < x < 1\). \[ Power = P(X \in C | H_1) = P(X > 1 - \alpha | H_1) = \int_{1-\alpha}^1 f_1(x) \, dx \] \[ Power = \int_{1-\alpha}^1 2x \, dx = [x^2]_{1-\alpha}^1 = 1^2 - (1-\alpha)^2 \] \[ Power = 1 - (1 - 2\alpha + \alpha^2) = 1 - 1 + 2\alpha - \alpha^2 = 2\alpha - \alpha^2 = \alpha(2-\alpha) \]
Thus, the power of the most powerful size \(\alpha\) test is \(\alpha(2-\alpha)\).
Quick Tip: For the Neyman-Pearson test, the critical region is always determined by the likelihood ratio. First, find the general form of the rejection region. Second, use the given size \(\alpha\) to find the specific critical value. Finally, calculate the probability of this critical region under the alternative hypothesis to find the power.


Question 9:

Suppose \(X \sim N(0,4)\) and \(Y \sim N(0,9)\) are independent random variables. Then the value of \(P(9X^2 + 4Y^2 < 6)\) is

  • (A) \(1 - e^{-1/4}\)
  • (B) \(1 - e^{-1/12}\)
  • (C) \(1 - e^{-1/6}\)
  • (D) \(1 - e^{-1/9}\)
Correct Answer: (B) \(1 - e^{-1/12}\)
View Solution




Step 1: Understanding the Concept:

We need to find the probability of an inequality involving the squares of two independent normal random variables. This suggests transforming the variables into standard normal variables and identifying the resulting distribution.

Recall that if \(Z \sim N(0,1)\), then \(Z^2 \sim \chi^2(1)\), which is the chi-squared distribution with 1 degree of freedom. A \(\chi^2(1)\) distribution is also a Gamma(\(\alpha=1/2, \beta=1/2\)) distribution. The sum of independent chi-squared variables is also a chi-squared variable.




Step 2: Standardizing the Variables:

Given \(X \sim N(0, 4)\), the standard deviation is \(\sigma_X = \sqrt{4} = 2\). Let \(Z_1 = \frac{X}{2}\). Then \(Z_1 \sim N(0,1)\).
Given \(Y \sim N(0, 9)\), the standard deviation is \(\sigma_Y = \sqrt{9} = 3\). Let \(Z_2 = \frac{Y}{3}\). Then \(Z_2 \sim N(0,1)\).
From these, we have \(X = 2Z_1\) and \(Y = 3Z_2\).




Step 3: Transforming the Inequality and Identifying the Distribution:

Substitute \(X\) and \(Y\) in the inequality: \[ P(9X^2 + 4Y^2 < 6) = P(9(2Z_1)^2 + 4(3Z_2)^2 < 6) \] \[ = P(9(4Z_1^2) + 4(9Z_2^2) < 6) \] \[ = P(36Z_1^2 + 36Z_2^2 < 6) \] \[ = P\left(Z_1^2 + Z_2^2 < \frac{6}{36}\right) = P\left(Z_1^2 + Z_2^2 < \frac{1}{6}\right) \]
Since \(Z_1\) and \(Z_2\) are independent standard normal variables, the sum of their squares, \(W = Z_1^2 + Z_2^2\), follows a chi-squared distribution with \(1+1=2\) degrees of freedom, i.e., \(W \sim \chi^2(2)\).

A \(\chi^2(2)\) distribution is equivalent to an Exponential distribution with rate \(\lambda = 1/2\), or mean \(\beta = 2\). Let's use the rate parameter \(\lambda=1/2\). The PDF of \(W\) is \(f(w) = \frac{1}{2}e^{-w/2}\) for \(w > 0\). The CDF is \(F(w) = P(W \le w) = 1 - e^{-w/2}\).




Step 4: Calculating the Probability:

We need to calculate \(P(W < 1/6)\). Using the CDF of the Exponential(rate=1/2) distribution: \[ P\left(W < \frac{1}{6}\right) = F\left(\frac{1}{6}\right) = 1 - e^{-(\frac{1}{6})/2} = 1 - e^{-1/12} \] Quick Tip: The sum of squares of \(n\) independent standard normal variables follows a chi-squared distribution with \(n\) degrees of freedom, \(\chi^2(n)\). Remember that a \(\chi^2(2)\) distribution is a special case and is identical to an Exponential distribution with mean 2. This can simplify probability calculations significantly.


Question 10:

Let \(X\) be a single sample from a continuous distribution with probability density function

where \(\theta > 0\) is an unknown parameter. For \(0 < \alpha < 0.05\), a \(100(1-\alpha)%\) confidence interval for \(\theta\) based on \(X\) is

  • (A) \(\left[ \frac{X}{1-\sqrt{\alpha/2}}, \frac{X}{1-\sqrt{1-\alpha/2}} \right]\)
  • (B) \(\left( \frac{X}{1-\sqrt{\alpha}}, \frac{X}{1-\sqrt{1-\alpha}} \right)\)
  • (C) \(\left( \left(1-\sqrt{1-\frac{\alpha}{2}}\right)X, \left(1-\sqrt{\frac{\alpha}{2}}\right)X \right)\)
  • (D) \(\left( \frac{\alpha}{2} X, \left(1-\frac{\alpha}{2}\right)X \right)\)
Correct Answer: (A) \(\left[ \frac{X}{1-\sqrt{\alpha/2}}, \frac{X}{1-\sqrt{1-\alpha/2}} \right]\)
View Solution




Step 1: Understanding the Concept:

We need to construct a confidence interval for the parameter \(\theta\) using the pivotal quantity method. A pivotal quantity is a function of the sample and the parameter whose distribution does not depend on the parameter.




Step 2: Finding a Pivotal Quantity:

First, let's find the cumulative distribution function (CDF) of \(X\). For \(0 < x < \theta\): \[ F(x; \theta) = P(X \le x) = \int_0^x \frac{2(\theta-t)}{\theta^2} \, dt = \frac{2}{\theta^2} \left[ \theta t - \frac{t^2}{2} \right]_0^x = \frac{2}{\theta^2} \left( \theta x - \frac{x^2}{2} \right) = \frac{2x}{\theta} - \frac{x^2}{\theta^2} \]
Let's check if there is a simpler way. Let \(Y = X/\theta\). We find the distribution of Y. For \(0 < y < 1\), the CDF of Y is: \[ F_Y(y) = P(Y \le y) = P(X/\theta \le y) = P(X \le y\theta) = F_X(y\theta) \] \[ F_Y(y) = \frac{2(y\theta)}{\theta} - \frac{(y\theta)^2}{\theta^2} = 2y - y^2 \]
Since the CDF \(F_Y(y) = 2y - y^2\) for \(0 < y < 1\) is free of \(\theta\), the random variable \(Q = X/\theta\) is a pivotal quantity.




Step 3: Constructing the Confidence Interval:

For a \(100(1-\alpha)%\) confidence interval, we need to find constants \(c_1\) and \(c_2\) such that: \[ P(c_1 < Q < c_2) = 1 - \alpha \]
where \(Q=X/\theta\). A standard way is to choose an equal-tailed interval, meaning: \(P(Q \le c_1) = \alpha/2\) and \(P(Q \ge c_2) = \alpha/2\), which implies \(P(Q \le c_2) = 1 - \alpha/2\).
We use the CDF of \(Q\), which is \(F_Q(q) = 2q - q^2\).

Find \(c_1\): \(F_Q(c_1) = \alpha/2 \implies 2c_1 - c_1^2 = \alpha/2\).
This is a quadratic equation for \(c_1\): \(c_1^2 - 2c_1 + \alpha/2 = 0\).
Using the quadratic formula, \(c_1 = \frac{2 \pm \sqrt{4 - 4(\alpha/2)}}{2} = 1 \pm \sqrt{1 - \alpha/2}\). Since \(c_1\) must be between 0 and 1, we choose the minus sign: \(c_1 = 1 - \sqrt{1 - \alpha/2}\).

Find \(c_2\): \(F_Q(c_2) = 1 - \alpha/2 \implies 2c_2 - c_2^2 = 1 - \alpha/2\).
This gives the equation \(c_2^2 - 2c_2 + (1 - \alpha/2) = 0\).
Factoring this, we get \((c_2 - 1)^2 - (\sqrt{\alpha/2})^2 = 0\)? No.
Let's rewrite it as \(c_2^2 - 2c_2 + 1 = \alpha/2 \implies (c_2-1)^2 = \alpha/2\).
So, \(c_2 - 1 = \pm \sqrt{\alpha/2}\). This gives \(c_2 = 1 \pm \sqrt{\alpha/2}\). Since \(c_2\) must be between 0 and 1, we choose the minus sign: \(c_2 = 1 - \sqrt{\alpha/2}\).

So, we have the probability statement: \[ P\left( 1 - \sqrt{1 - \alpha/2} < \frac{X}{\theta} < 1 - \sqrt{\alpha/2} \right) = 1 - \alpha \]



Step 4: Inverting the Inequality to find the Interval for \(\theta\):

We need to isolate \(\theta\) in the middle of the inequality.
The left inequality: \(1 - \sqrt{1 - \alpha/2} < X/\theta \implies \theta < \frac{X}{1 - \sqrt{1 - \alpha/2}}\).
The right inequality: \(X/\theta < 1 - \sqrt{\alpha/2} \implies \theta > \frac{X}{1 - \sqrt{\alpha/2}}\).
Combining these, we get: \[ \frac{X}{1 - \sqrt{\alpha/2}} < \theta < \frac{X}{1 - \sqrt{1 - \alpha/2}} \]
The \(100(1-\alpha)%\) confidence interval for \(\theta\) is: \[ \left( \frac{X}{1 - \sqrt{\alpha/2}}, \frac{X}{1 - \sqrt{1 - \alpha/2}} \right) \] Quick Tip: The pivotal quantity method is a standard technique for constructing confidence intervals. The key steps are: 1. Find the CDF of the random variable \(X\). 2. Define a pivotal quantity \(Q = g(X, \theta)\) whose distribution is independent of \(\theta\). Often, \(Q = F(X; \theta)\) which is \(U(0,1)\), or a simple transformation like \(X/\theta\). 3. Find quantiles \(c_1, c_2\) for the pivotal quantity's distribution. 4. Invert the probability statement \(P(c_1 < Q < c_2) = 1-\alpha\) to get an interval for \(\theta\).


Question 11:

Let \(f: \mathbb{R} \to \mathbb{R}\) be given by \(f(x) = x + \pi \cos x\). Then the number of solutions of the equation \(f(x) = 0\) is

  • (A) 1
  • (B) 2
  • (C) 3
  • (D) 4
Correct Answer: (C) 3
View Solution




Step 1: Understanding the Concept:

We need to find the number of roots of the equation \(x + \pi \cos x = 0\), which is equivalent to finding the number of intersection points between the graphs of \(y = x\) and \(y = -\pi \cos x\). A more robust method is to analyze the function \(f(x)\) using calculus.




Step 2: Analyzing the function \(f(x)\) using its derivative:

Let's find the derivative of \(f(x)\) to understand its increasing/decreasing nature. \[ f(x) = x + \pi \cos x \] \[ f'(x) = 1 - \pi \sin x \]
The critical points occur when \(f'(x) = 0\), which means \(1 - \pi \sin x = 0\), or \(\sin x = 1/\pi\).
Since \( \pi \approx 3.14 \), we have \(0 < 1/\pi < 1\). This means the equation \(\sin x = 1/\pi\) has infinitely many solutions. These solutions correspond to the local maxima and minima of \(f(x)\).




Step 3: Finding the intervals of increase/decrease and evaluating \(f(x)\) at critical points:

- \(f'(x) > 0\) when \(1 - \pi \sin x > 0 \implies \sin x < 1/\pi\). In these intervals, \(f(x)\) is increasing.
- \(f'(x) < 0\) when \(1 - \pi \sin x < 0 \implies \sin x > 1/\pi\). In these intervals, \(f(x)\) is decreasing.

Let's evaluate \(f(x)\) at some key points to locate the roots.

\(f(0) = 0 + \pi \cos(0) = \pi > 0\)
\(f(\pi/2) = \pi/2 + \pi \cos(\pi/2) = \pi/2 > 0\)
\(f(-\pi/2) = -\pi/2 + \pi \cos(-\pi/2) = -\pi/2 < 0\).
Since \(f\) is continuous and \(f(-\pi/2) < 0\) and \(f(0) > 0\), by the Intermediate Value Theorem (IVT), there must be at least one root in \((-\pi/2, 0)\).

\(f(\pi) = \pi + \pi \cos(\pi) = \pi - \pi = 0\). So, \(x = \pi\) is a root.

\(f(-\pi) = -\pi + \pi \cos(-\pi) = -\pi - \pi = -2\pi < 0\).
\(f(-3\pi/2) = -3\pi/2 + \pi \cos(-3\pi/2) = -3\pi/2 < 0\).



Let's summarize the roots found so far:
1. One root in \((-\pi/2, 0)\).
2. One root at \(x = \pi\).
3. One root in \((\pi/2, x_0) \subset (\pi/2, \pi)\).

Let's check the interval \([-\pi, -\pi/2]\). \(f(-\pi) = -2\pi < 0\). \(f(-\pi/2) = -\pi/2 < 0\).
Let's analyze \(f'(x)\) in this interval. \(\sin x\) goes from 0 to -1. So \(\sin x\) is always negative, which means \(\sin x < 1/\pi\).
Therefore, \(f'(x) = 1 - \pi \sin x > 0\) for all \(x \in (-\pi, -\pi/2)\).
This means \(f(x)\) is strictly increasing from \(f(-\pi)=-2\pi\) to \(f(-\pi/2)=-\pi/2\). Since the function values remain negative, there are no roots in this interval.

So, we have found exactly three roots in the interval \([-\pi, \pi]\). Since any root must be in this interval, there are a total of 3 solutions. Quick Tip: To find the number of roots of an equation \(g(x) = h(x)\), it is often effective to define a function \(f(x) = g(x) - h(x)\) and analyze its properties. Use the derivative \(f'(x)\) to find intervals of increase/decrease and local extrema. Then use the Intermediate Value Theorem by evaluating \(f(x)\) at the endpoints of these intervals and at the extrema to count the number of times \(f(x)\) crosses the x-axis.


Question 12:

A fair die is thrown three times independently. The probability that 4 is the maximum value that appears among these throws is equal to

  • (A) \( \frac{8}{27} \)
  • (B) \( \frac{1}{216} \)
  • (C) \( \frac{37}{216} \)
  • (D) \( \frac{1}{2} \)
Correct Answer: (C) \( \frac{37}{216} \)
View Solution




Step 1: Understanding the Concept:

Let \(X_1, X_2, X_3\) be the outcomes of the three throws. We want to find the probability that the maximum of these three values is exactly 4. Let \(M = \max(X_1, X_2, X_3)\). We want to calculate \(P(M=4)\).




Step 2: Key Formula or Approach:

The event \(\{M=4\}\) is difficult to count directly. A standard technique is to use the complementary event or a related event.
The event \(\{M=4\}\) can be expressed as the difference between two events: \(\{ All outcomes are less than or equal to 4 \}\) MINUS \(\{ All outcomes are less than or equal to 3 \}\).
In symbols: \(P(M=4) = P(M \le 4) - P(M \le 3)\).




Step 3: Detailed Explanation:

The total number of possible outcomes for three throws is \(6 \times 6 \times 6 = 216\).

Calculate \(P(M \le 4)\):

The event \(M \le 4\) means that for each throw, the outcome must be in the set \(\{1, 2, 3, 4\}\).
The probability of getting a number \(\le 4\) in a single throw is \(P(X_i \le 4) = 4/6 = 2/3\).
Since the throws are independent, the probability that all three throws are \(\le 4\) is:
\[ P(M \le 4) = P(X_1 \le 4 and X_2 \le 4 and X_3 \le 4) = \left(\frac{4}{6}\right)^3 = \left(\frac{2}{3}\right)^3 = \frac{8}{27} \]

Calculate \(P(M \le 3)\):

The event \(M \le 3\) means that for each throw, the outcome must be in the set \(\{1, 2, 3\}\).
The probability of getting a number \(\le 3\) in a single throw is \(P(X_i \le 3) = 3/6 = 1/2\).
Since the throws are independent, the probability that all three throws are \(\le 3\) is:
\[ P(M \le 3) = P(X_1 \le 3 and X_2 \le 3 and X_3 \le 3) = \left(\frac{3}{6}\right)^3 = \left(\frac{1}{2}\right)^3 = \frac{1}{8} \]

Now, we can find \(P(M=4)\): \[ P(M=4) = P(M \le 4) - P(M \le 3) = \frac{8}{27} - \frac{1}{8} \]
To subtract the fractions, we find a common denominator, which is \(27 \times 8 = 216\). \[ P(M=4) = \frac{8 \times 8}{27 \times 8} - \frac{1 \times 27}{8 \times 27} = \frac{64}{216} - \frac{27}{216} = \frac{64 - 27}{216} = \frac{37}{216} \]



Step 4: Final Answer:

The probability that 4 is the maximum value is \(37/216\).
Quick Tip: For problems involving the maximum (or minimum) of several independent random variables, the CDF approach is often the easiest. The probability \(P(\max(X_i) \le k)\) is equal to \(P(all X_i \le k)\), which, by independence, is \(\prod P(X_i \le k)\). Then, \(P(\max(X_i) = k) = P(\max(X_i) \le k) - P(\max(X_i) \le k-1)\).


Question 13:

Let \(\mathcal{A}\) be an \(n \times n\) matrix. Which of the following statements is NOT necessarily true?

  • (A) If \(rank(\mathcal{A}^5) = rank(\mathcal{A}^6)\), then \(rank(\mathcal{A}^6) = rank(\mathcal{A}^7)\)
  • (B) If \(rank(\mathcal{A}) = n\), then it is possible to obtain a singular matrix by suitably changing a single entry of \(\mathcal{A}\)
  • (C) If \(rank(\mathcal{A}) = n\), then \(rank(\mathcal{A} + \mathcal{A}^T) \ge \frac{n}{2}\)
  • (D) If \(rank(\mathcal{A}) < n\), then it is possible to obtain a nonsingular matrix by suitably changing \(n - rank(\mathcal{A})\) entries of \(\mathcal{A}\)
Correct Answer: (C) If \(\text{rank}(\mathcal{A}) = n\), then \(\text{rank}(\mathcal{A} + \mathcal{A}^T) \ge \frac{n}{2}\)
View Solution



We analyze each statement to determine its validity. We are looking for the statement that is not always true.

Step 1: Analyzing Statement (A):

This statement concerns the sequence of ranks of powers of a matrix. The column space (image) of \(\mathcal{A}^{k+1}\) is a subspace of the column space of \(\mathcal{A}^k\), i.e., \(Im(\mathcal{A}^{k+1}) \subseteq Im(\mathcal{A}^k)\). This implies the sequence of ranks is non-increasing: \(rank(\mathcal{A}) \ge rank(\mathcal{A}^2) \ge \dots\). Once the rank stabilizes, i.e., \(rank(\mathcal{A}^k) = rank(\mathcal{A}^{k+1})\) for some \(k\), it implies the corresponding subspaces are equal: \(Im(\mathcal{A}^k) = Im(\mathcal{A}^{k+1})\). Applying the linear transformation \(\mathcal{A}\) to both sides gives \(\mathcal{A}(Im(\mathcal{A}^k)) = \mathcal{A}(Im(\mathcal{A}^{k+1}))\), which means \(Im(\mathcal{A}^{k+1}) = Im(\mathcal{A}^{k+2})\). This shows that \(rank(\mathcal{A}^{k+1}) = rank(\mathcal{A}^{k+2})\). This stability continues for all higher powers. Therefore, if \(rank(\mathcal{A}^5) = rank(\mathcal{A}^6)\), the sequence has stabilized, and it must be that \(rank(\mathcal{A}^6) = rank(\mathcal{A}^7)\). Statement (A) is necessarily true.




Step 2: Analyzing Statement (B):

If \(rank(\mathcal{A}) = n\), the matrix is nonsingular, and its determinant \(\det(\mathcal{A})\) is non-zero. The determinant can be expressed using cofactor expansion along any row or column. For instance, along the \(i\)-th row: \(\det(\mathcal{A}) = \sum_{j=1}^n a_{ij} C_{ij}\), where \(C_{ij}\) is the \((i, j)\)-cofactor. If we change a single entry \(a_{ij}\) to a variable \(x\), the new determinant becomes a linear function of \(x\): \(\det(\mathcal{A}(x)) = x C_{ij} + (terms not involving a_{ij})\). Since \(\mathcal{A}\) is nonsingular, at least one cofactor must be non-zero. Thus, \(\det(\mathcal{A}(x))\) is a non-constant linear function of \(x\), which has exactly one root. By choosing \(x\) to be this root, we can make the determinant zero, and the matrix becomes singular. Statement (B) is necessarily true.




Step 3: Analyzing Statement (C):

If \(rank(\mathcal{A}) = n\), the matrix \(\mathcal{A}\) is invertible. We need to check if \(rank(\mathcal{A} + \mathcal{A}^T) \ge \frac{n}{2}\) must hold. Let's look for a counterexample.
Consider the case where \(\mathcal{A}\) is a skew-symmetric matrix (\(\mathcal{A}^T = -\mathcal{A}\)). For a skew-symmetric matrix to be nonsingular, its dimension \(n\) must be even.
Let \(n=2\) and consider the matrix . \(\det(\mathcal{A}) = 0 - (-1) = 1 \ne 0\), so \(rank(\mathcal{A}) = 2 = n\).
Now, let's compute \(\mathcal{A} + \mathcal{A}^T\):

The rank of the zero matrix is 0.
The statement requires \(rank(\mathcal{A} + \mathcal{A}^T) \ge n/2\), which for our example is \(0 \ge 2/2 = 1\). This is false.
Since we have found a valid counterexample, statement (C) is NOT necessarily true.

Step 4: Analyzing Statement (D):

This statement claims that a singular matrix of rank \(r\) can be made nonsingular by changing \(n-r\) entries. This is a known result from matrix theory. The rank deficiency is \(d = n-r\). It is possible to increase the rank of a matrix by 1 by changing a single, suitably chosen entry. By repeating this process \(d\) times, we can increase the rank from \(r\) to \(r+d = n\). Thus, it is always possible to make the matrix nonsingular by changing \(d=n-r\) entries. Statement (D) is necessarily true.

Final Answer:

Statements (A), (B), and (D) are necessarily true. Statement (C) is not necessarily true, as demonstrated by the counterexample of a nonsingular skew-symmetric matrix. Quick Tip: When asked to find a statement that is "not necessarily true" in linear algebra, actively search for counterexamples. Special types of matrices like symmetric, skew-symmetric, diagonal, or nilpotent matrices are often good candidates for constructing these counterexamples.


Question 14:

Let \(V\) be a subspace of \(\mathbb{R}^{10}\). Suppose \(\mathcal{A}\) is a \(10 \times 10\) matrix with real entries. Let \(\mathcal{A}^k(V) = \{\mathcal{A}^k\mathbf{x} : \mathbf{x} \in V\}\) for \(k \ge 1\) and \(\mathcal{A}(V) = \mathcal{A}^1(V)\). Which one of the following statements is NOT true?

  • (A) If \(\mathcal{A}\) is nonsingular, then \(\dim(V) = \dim(\mathcal{A}(V))\) necessarily holds
  • (B) It is possible that \(\mathcal{A}\) is singular and \(\dim(V) = \dim(\mathcal{A}(V))\)
  • (C) If \(rank(\mathcal{A}) = 8\), then \(\dim(\mathcal{A}(V)) \ge \dim(V) - 2\) necessarily holds
  • (D) If \(\dim(V) = \dim(\mathcal{A}(V)) = \dim(\mathcal{A}^2(V)) = \dots = \dim(\mathcal{A}^5(V))\), then \(\dim(\mathcal{A}^6(V)) = \dim(V)\) necessarily holds
Correct Answer: (D) If \(\dim(V) = \dim(\mathcal{A}(V)) = \dim(\mathcal{A}^2(V)) = \dots = \dim(\mathcal{A}^5(V))\), then \(\dim(\mathcal{A}^6(V)) = \dim(V)\) necessarily holds
View Solution



Let \(T: \mathbb{R}^{10} \to \mathbb{R}^{10}\) be the linear transformation given by \(T(\mathbf{x}) = \mathcal{A}\mathbf{x}\). Let \(T_V\) be the restriction of \(T\) to the subspace \(V\). The image of this restricted map is \(\mathcal{A}(V)\).

(A) If \(\mathcal{A}\) is nonsingular, then \(\dim(V) = \dim(\mathcal{A}(V))\) necessarily holds.
If \(\mathcal{A}\) is nonsingular, then \(\ker(\mathcal{A}) = \{\mathbf{0}\}\). The kernel of the restricted map \(T_V\) is \(\ker(T_V) = \{\mathbf{x} \in V : \mathcal{A}\mathbf{x} = \mathbf{0}\} = V \cap \ker(\mathcal{A}) = V \cap \{\mathbf{0}\} = \{\mathbf{0}\}\). By the Rank-Nullity Theorem for \(T_V\), we have \(\dim(V) = \dim(\ker(T_V)) + \dim(Im(T_V))\). This gives \(\dim(V) = 0 + \dim(\mathcal{A}(V))\). Thus, the statement is true.

(B) It is possible that \(\mathcal{A}\) is singular and \(\dim(V) = \dim(\mathcal{A}(V))\).
For \(\dim(V) = \dim(\mathcal{A}(V))\) to hold, we need \(\dim(\ker(T_V)) = 0\), which means \(V \cap \ker(\mathcal{A}) = \{\mathbf{0}\}\). Since \(\mathcal{A}\) is singular, \(\ker(\mathcal{A})\) is a non-trivial subspace. We can choose a subspace \(V\) that is complementary to a subspace containing \(\ker(\mathcal{A})\) or simply has a trivial intersection with it. For example, in \(\mathbb{R}^2\), let . \(\ker(\mathcal{A})\) is the y-axis. Let \(V\) be the x-axis. Then \(V \cap \ker(\mathcal{A}) = \{\mathbf{0}\}\), and \(\dim(V)=1\). \(\mathcal{A}(V)\) is also the x-axis, so \(\dim(\mathcal{A}(V))=1\). Thus, the statement is true.

(C) If \(rank(\mathcal{A}) = 8\), then \(\dim(\mathcal{A}(V)) \ge \dim(V) - 2\) necessarily holds.

By the Rank-Nullity Theorem for \(\mathcal{A}\), \(\dim(\mathbb{R}^{10}) = rank(\mathcal{A}) + \dim(\ker(\mathcal{A}))\), so \(10 = 8 + \dim(\ker(\mathcal{A}))\), which implies \(\dim(\ker(\mathcal{A})) = 2\).

For the restricted map \(T_V\), \(\dim(\mathcal{A}(V)) = \dim(V) - \dim(\ker(T_V)) = \dim(V) - \dim(V \cap \ker(\mathcal{A}))\).

Since \(\dim(V \cap \ker(\mathcal{A})) \le \dim(\ker(\mathcal{A}))\), we have \(\dim(V \cap \ker(\mathcal{A})) \le 2\).

Therefore, \(\dim(\mathcal{A}(V)) \ge \dim(V) - 2\). The statement is true.

(D) If \(\dim(V) = \dim(\mathcal{A}(V)) = \dots = \dim(\mathcal{A}^5(V))\), then \(\dim(\mathcal{A}^6(V)) = \dim(V))\) necessarily holds.
The general relation is \(\dim(\mathcal{A}^{k+1}(V)) \le \dim(\mathcal{A}^k(V))\). The given condition implies that the dimension is constant for \(k=0\) to \(k=5\). Let's construct a counterexample.

Consider \(\mathbb{R}^{10}\) with the standard basis \(\{\mathbf{e}_1, \dots, \mathbf{e}_{10}\}\). Define a linear transformation \(\mathcal{A}\) as the nilpotent shift operator: \(\mathcal{A}\mathbf{e}_i = \mathbf{e}_{i-1}\) for \(i=2, \dots, 10\) and \(\mathcal{A}\mathbf{e}_1 = \mathbf{0}\).

The kernel of \(\mathcal{A}\) is \(\ker(\mathcal{A}) = span\{\mathbf{e}_1\}\).
Let \(V\) be the one-dimensional subspace spanned by \(\mathbf{e}_{10}\), i.e., \(V = span\{\mathbf{e}_{10}\}\). \(\dim(V)=1\).

Let's compute the dimensions of the images:

\(\mathcal{A}(V) = \mathcal{A}(span\{\mathbf{e}_{10}\}) = span\{\mathbf{e}_9\}\). So \(\dim(\mathcal{A}(V))=1\).
\(\mathcal{A}^2(V) = \mathcal{A}(span\{\mathbf{e}_9\}) = span\{\mathbf{e}_8\}\). So \(\dim(\mathcal{A}^2(V))=1\).
...
\(\mathcal{A}^5(V) = span\{\mathbf{e}_5\}\). So \(\dim(\mathcal{A}^5(V))=1\).

The premise of the statement holds: \(\dim(V) = \dim(\mathcal{A}(V)) = \dots = \dim(\mathcal{A}^5(V)) = 1\).

Now let's check the conclusion: \(\mathcal{A}^6(V) = span\{\mathbf{e}_4\}\), so \(\dim(\mathcal{A}^6(V))=1\).
Wait, my counterexample needs to be more careful. The shift operator is \(A^k e_i = e_{i-k}\).
Let \(V = span\{e_7\}\). \(\dim(V)=1\). \(\dim(A(V)) = \dim(span\{e_6\})=1\).

... \(\dim(A^5(V)) = \dim(span\{e_2\})=1\).
The premise \(\dim(V) = \dots = \dim(A^5(V)) = 1\) is satisfied.
Now, \(\dim(A^6(V)) = \dim(span\{e_1\})=1\). The conclusion \(\dim(A^6(V))=\dim(V)\) holds here.
But what about \(\dim(A^7(V))\)? \(A^7(V) = A(span\{e_1\}) = \{0\}\), so \(\dim(A^7(V))=0\).

The statement only asks about \(A^6(V)\). Let's re-examine my first counterexample.
Let \(V = span\{\mathbf{e}_6\}\) in \(\mathbb{R}^{10}\). \(\dim(V)=1\). \(\dim(A(V)) = \dim(span\{\mathbf{e}_5\})=1\).
\(\dim(A^2(V)) = \dim(span\{\mathbf{e}_4\})=1\).
\(\dim(A^3(V)) = \dim(span\{\mathbf{e}_3\})=1\).
\(\dim(A^4(V)) = \dim(span\{\mathbf{e}_2\})=1\).
\(\dim(A^5(V)) = \dim(span\{\mathbf{e}_1\})=1\).

The premise is satisfied. Now for the conclusion: \(\mathcal{A}^6(V) = \mathcal{A}(\mathcal{A}^5(V)) = \mathcal{A}(span\{\mathbf{e}_1\}) = span\{\mathcal{A}\mathbf{e}_1\} = span\{\mathbf{0}\} = \{\mathbf{0}\}\).

So, \(\dim(\mathcal{A}^6(V)) = 0\).

However, \(\dim(V) = 1\).

Therefore, \(\dim(\mathcal{A}^6(V)) \ne \dim(V)\). The statement does not necessarily hold.
This statement is NOT true. Quick Tip: When analyzing properties of linear transformations and subspaces, the Rank-Nullity Theorem (\(\dim(W) = \dim(\ker(T_W)) + \dim(Im(T_W))\)) is a fundamental tool. For statements about sequences of transformations, like powers of a matrix, nilpotent matrices (like the shift operator) are excellent sources of counterexamples.


Question 15:

A function \(f:(0,1) \to \mathbb{R}\) is said to have property \(\mathcal{I}\) if, for any \(0 < x_1 < x_2 < 1\) and for any \(c\) between \(f(x_1)\) and \(f(x_2)\), there exists \(y \in [x_1, x_2]\) such that \(f(y)=c\). Consider the following statements:

[(I)] If \(g:(0,1) \to \mathbb{R}\) satisfies property \(\mathcal{I}\), then \(g\) is necessarily continuous.
[(II)] If \(h:(0,1) \to \mathbb{R}\) is differentiable, then \(h'\) necessarily satisfies property \(\mathcal{I}\).

Then

  • (A) (I) is correct and (II) is incorrect
  • (B) (I) is incorrect and (II) is correct
  • (C) Both (I) and (II) are correct
  • (D) Both (I) and (II) are incorrect
Correct Answer: (B) (I) is incorrect and (II) is correct
View Solution




Step 1: Understanding Property \(\mathcal{I}\):

Property \(\mathcal{I}\) is the Intermediate Value Property (IVP). The Intermediate Value Theorem states that all continuous functions have this property. The question here is about the converse and about whether derivatives have this property.

Step 2: Analyzing Statement (I):

This statement claims that if a function has the Intermediate Value Property, then it must be continuous. This is a well-known false statement in real analysis. A function can satisfy the IVP without being continuous.
A classic counterexample is the derivative of certain functions. For instance, consider the function \(H(x) = x^2 \sin(1/x)\) for \(x \ne 0\) and \(H(0) = 0\). This function is differentiable on \(\mathbb{R}\). Its derivative is:

The derivative \(H'(x)\) is not continuous at \(x=0\) because \(\cos(1/x)\) oscillates and does not approach a limit as \(x \to 0\). However, by Darboux's Theorem (see Statement II), \(H'(x)\) must have the Intermediate Value Property.
If we consider the function \(g(x) = H'(x)\) on the interval \((0,1)\), this function \(g\) is not continuous (it has discontinuities at points \(1/(k\pi)\) if we extend it, but more importantly it demonstrates the principle). The key point is that a function can have the IVP without being continuous.
Therefore, Statement (I) is incorrect.

Step 3: Analyzing Statement (II):

This statement claims that the derivative of any differentiable function necessarily satisfies property \(\mathcal{I}\). This is the statement of Darboux's Theorem. Darboux's Theorem states that if a function \(h\) is differentiable on an interval \([a,b]\), then its derivative \(h'\) has the Intermediate Value Property on \([a,b]\). This means that for any value \(c\) between \(h'(a)\) and \(h'(b)\), there exists some \(y \in (a,b)\) such that \(h'(y) = c\). This theorem extends to any subinterval of the domain of differentiability.
Therefore, Statement (II) is correct.

Step 4: Final Answer:

Statement (I) is incorrect, and Statement (II) is correct. This corresponds to option (B).
Quick Tip: Remember the relationship between continuity, differentiability, and the Intermediate Value Property (IVP): Continuity \(\implies\) IVP (Intermediate Value Theorem) IVP \(\not\implies\) Continuity (e.g., derivative of \(x^2\sin(1/x)\)) Differentiability of \(h\) \(\implies\) IVP for \(h'\) (Darboux's Theorem)


Question 16:

Suppose \(f\) is a polynomial of degree \(n\) with real coefficients, and \(\mathcal{A}\) is an \(n \times n\) matrix with real entries satisfying \(f(\mathcal{A}) = 0\). Consider the following statements:

[(I)] If \(f(0) \ne 0\), then \(\mathcal{A}\) is necessarily nonsingular.
[(II)] If \(f(0) = 0\), then \(\mathcal{A}\) is necessarily singular.

Then

  • (A) (I) is correct and (II) is incorrect
  • (B) (I) is incorrect and (II) is correct
  • (C) Both (I) and (II) are correct
  • (D) Both (I) and (II) are incorrect
  • (I) If \(f(0) \ne 0\), then 0 is not a root of \(f(x)\), so 0 cannot be an eigenvalue of \(\mathcal{A}\). Thus \(\mathcal{A}\) is nonsingular.
Correct Answer: (A) (I) is correct and (II) is incorrect
View Solution




Step 1: Understanding the Concept:

The equation \(f(\mathcal{A}) = 0\) means that \(\mathcal{A}\) is a root of the polynomial \(f(x)\). This implies that the minimal polynomial of \(\mathcal{A}\), denoted \(m_{\mathcal{A}}(x)\), must divide \(f(x)\). A matrix \(\mathcal{A}\) is nonsingular (invertible) if and only if 0 is not an eigenvalue of \(\mathcal{A}\). The eigenvalues of \(\mathcal{A}\) are roots of its minimal polynomial.




Step 2: Analyzing Statement (I):

We are given \(f(0) \ne 0\). Let the polynomial be \(f(x) = c_n x^n + \dots + c_1 x + c_0\). Then \(f(0) = c_0 \ne 0\).
The condition \(f(\mathcal{A}) = 0\) means \(c_n \mathcal{A}^n + \dots + c_1 \mathcal{A} + c_0 I = 0\), where \(I\) is the identity matrix.
We can rearrange this equation: \[ c_n \mathcal{A}^n + \dots + c_1 \mathcal{A} = -c_0 I \] \[ \mathcal{A} (c_n \mathcal{A}^{n-1} + \dots + c_1 I) = -c_0 I \]
Since \(c_0 \ne 0\), we can divide by \(-c_0\): \[ \mathcal{A} \left( -\frac{1}{c_0} (c_n \mathcal{A}^{n-1} + \dots + c_1 I) \right) = I \]
This shows that \(\mathcal{A}\) has a right inverse. For a square matrix, having a right inverse implies it is invertible (nonsingular).
Alternatively, assume \(\mathcal{A}\) is singular. Then it has an eigenvalue \(\lambda = 0\). Let \(\mathbf{v} \ne \mathbf{0}\) be a corresponding eigenvector, so \(\mathcal{A}\mathbf{v} = 0\mathbf{v} = \mathbf{0}\).
Applying \(f(\mathcal{A})\) to \(\mathbf{v}\): \(f(\mathcal{A})\mathbf{v} = (c_n \mathcal{A}^n + \dots + c_1 \mathcal{A} + c_0 I)\mathbf{v} = c_n \mathcal{A}^n\mathbf{v} + \dots + c_1 \mathcal{A}\mathbf{v} + c_0 I\mathbf{v}\).
Since \(\mathcal{A}^k\mathbf{v} = \mathbf{0}\) for \(k \ge 1\), this simplifies to \(f(\mathcal{A})\mathbf{v} = c_0 \mathbf{v}\).
We are given \(f(\mathcal{A}) = 0\), so \(0 \cdot \mathbf{v} = \mathbf{0}\).
This gives \(c_0 \mathbf{v} = \mathbf{0}\). Since \(\mathbf{v} \ne \mathbf{0}\), we must have \(c_0=0\), i.e., \(f(0)=0\).
This contradicts the given condition \(f(0) \ne 0\). Therefore, the assumption that \(\mathcal{A}\) is singular must be false.
Thus, Statement (I) is correct.




Step 3: Analyzing Statement (II):

We are given \(f(0) = 0\). This means the constant term of \(f(x)\) is zero, so \(x\) is a factor of \(f(x)\). We can write \(f(x) = x \cdot g(x)\) for some polynomial \(g(x)\).
The condition \(f(\mathcal{A})=0\) becomes \(\mathcal{A} \cdot g(\mathcal{A}) = 0\).
Taking the determinant of both sides: \(\det(\mathcal{A} \cdot g(\mathcal{A})) = \det(0)\), which means \(\det(\mathcal{A}) \cdot \det(g(\mathcal{A})) = 0\).
This implies that either \(\det(\mathcal{A}) = 0\) or \(\det(g(\mathcal{A})) = 0\).
The statement says \(\mathcal{A}\) is *necessarily* singular, which means \(\det(\mathcal{A})\) must be 0.
Is it possible for \(\mathcal{A}\) to be nonsingular? If \(\mathcal{A}\) is nonsingular, then \(\det(\mathcal{A}) \ne 0\), which would force \(\det(g(\mathcal{A}))=0\).
Let's construct a counterexample. We need a nonsingular matrix \(\mathcal{A}\) and a polynomial \(f(x)\) of degree \(n\) such that \(f(0)=0\) and \(f(\mathcal{A})=0\).
Let \(n=2\) and let \(\mathcal{A} = I_2 = \begin{pmatrix} 1 & 0
0 & 1 \end{pmatrix}\). This matrix is nonsingular.
We need a polynomial \(f(x)\) of degree 2 such that \(f(0)=0\) and \(f(I_2)=0\).
The minimal polynomial of \(I_2\) is \(m(x)=x-1\). Since \(f(I_2)=0\), \(m(x)\) must divide \(f(x)\). So, \((x-1)\) must be a factor of \(f(x)\).
The condition \(f(0)=0\) means \(x\) must be a factor of \(f(x)\).
Let's choose \(f(x) = x(x-1) = x^2-x\). This polynomial has degree \(n=2\).
It satisfies \(f(0)=0\).
It also satisfies \(f(\mathcal{A}) = f(I_2) = I_2^2 - I_2 = I_2 - I_2 = 0\).
So, we have found a nonsingular matrix \(\mathcal{A}=I_2\) that satisfies the conditions of the statement. This means \(\mathcal{A}\) is not necessarily singular.
Therefore, Statement (II) is incorrect.




Step 4: Final Answer:

Statement (I) is correct and Statement (II) is incorrect. This corresponds to option (A). Quick Tip: A matrix \(\mathcal{A}\) is singular if and only if 0 is one of its eigenvalues. The eigenvalues of \(\mathcal{A}\) are roots of its minimal polynomial \(m_{\mathcal{A}}(x)\). Since \(f(\mathcal{A})=0\), \(m_{\mathcal{A}}(x)\) must divide \(f(x)\). Therefore, all eigenvalues of \(\mathcal{A}\) must be roots of \(f(x)\). (I) If \(f(0) \ne 0\), then 0 is not a root of \(f(x)\), so 0 cannot be an eigenvalue of \(\mathcal{A}\). Thus \(\mathcal{A}\) is nonsingular. (II) If \(f(0) = 0\), then 0 is a root of \(f(x)\). This allows for the possibility that 0 is an eigenvalue, but does not require it, because the minimal polynomial \(m_{\mathcal{A}}(x)\) might be a factor of \(f(x)\) that does not include the factor \(x\).


Question 17:

Let \(X_1\) and \(X_2\) be i.i.d. \(N(0, \sigma^2)\) random variables. Define \(Z_1 = X_1 + X_2\) and \(Z_2 = X_1 - X_2\). Then which one of the following statements is NOT correct?

  • (A) \(Z_1\) and \(Z_2\) are independently distributed
  • (B) \(Z_1\) and \(Z_2\) are identically distributed
  • (C) \(P\left(\left|\frac{Z_1}{Z_2}\right| < 1\right) = 0.5\)
  • (D) \(\frac{Z_1}{Z_2}\) and \(Z_1^2 + Z_2^2\) are independently distributed
Correct Answer: (C) \(P\left(\left|\frac{Z_1}{Z_2}\right| < 1\right) = 0.5\)
View Solution



Let's analyze the properties of \(Z_1\) and \(Z_2\). This is a linear transformation of a bivariate normal vector \((X_1, X_2)\).

Step 1: Find the distribution of \((Z_1, Z_2)\).

Since \(X_1\) and \(X_2\) are independent normal variables, the vector \(\mathbf{X} = (X_1, X_2)^T\) has a bivariate normal distribution with mean vector \(\boldsymbol{\mu} = (0, 0)^T\) and covariance matrix .
The transformation is  \( = A\mathbf{X}\).
Since this is a linear transformation of a normal vector, \(\mathbf{Z}\) is also bivariate normal.
The mean of \(\mathbf{Z}\) is \(E[\mathbf{Z}] = A E[\mathbf{X}] = A\boldsymbol{\mu} = \mathbf{0}\).
The covariance matrix of \(\mathbf{Z}\) is \(\Sigma_Z = A \Sigma A^T\).

From this covariance matrix, we can see:

Var(\(Z_1\)) = \(2\sigma^2\)
Var(\(Z_2\)) = \(2\sigma^2\)
Cov(\(Z_1, Z_2\)) = 0

So, \(Z_1 \sim N(0, 2\sigma^2)\) and \(Z_2 \sim N(0, 2\sigma^2)\).

Step 2: Evaluate the statements.

(A) \(Z_1\) and \(Z_2\) are independently distributed.
Since \((Z_1, Z_2)\) have a bivariate normal distribution and their covariance is 0, they are independent. This statement is correct.

(B) \(Z_1\) and \(Z_2\) are identically distributed.
Both \(Z_1\) and \(Z_2\) follow the same distribution, \(N(0, 2\sigma^2)\). So they are identically distributed. This statement is correct.

(C) \(P\left(\left|\frac{Z_1}{Z_2}\right| < 1\right) = 0.5\).
Let \(U = Z_1/\sqrt{2\sigma^2}\) and \(V = Z_2/\sqrt{2\sigma^2}\). Then \(U, V\) are i.i.d. \(N(0,1)\).
The ratio \(Z_1/Z_2 = U/V\) follows a standard Cauchy distribution. Let \(T = Z_1/Z_2\).
The condition is \(|T| < 1\), which means \(-1 < T < 1\).
The CDF of a standard Cauchy distribution is \(F(t) = \frac{1}{\pi}\arctan(t) + \frac{1}{2}\). \[ P(-1 < T < 1) = F(1) - F(-1) = \left(\frac{1}{\pi}\arctan(1) + \frac{1}{2}\right) - \left(\frac{1}{\pi}\arctan(-1) + \frac{1}{2}\right) \] \[ = \frac{1}{\pi}\left(\frac{\pi}{4}\right) - \frac{1}{\pi}\left(-\frac{\pi}{4}\right) = \frac{1}{4} - \left(-\frac{1}{4}\right) = \frac{1}{2} = 0.5 \] Quick Tip: When dealing with linear combinations of independent normal variables, the result is also normal. Its mean and variance can be calculated using standard formulas. For two such variables, if their covariance is zero, they are independent. This is a special property of the multivariate normal distribution.


Question 18:

Consider a circle \(C\) with unit radius and center at \(A=(0,0)\). Let \(B=(1,0)\). Suppose \(\Theta \sim U(0, \pi)\) and \(D=(\cos\Theta, \sin\Theta)\). Note that the angle \(\angle DAB = \Theta\). Then the expected area of the triangle ABD is

  • (A) \( \frac{1}{\pi} \)
  • (B) \( \frac{2}{\pi} \)
  • (C) \( \frac{1}{2\pi} \)
  • (D) 1
Correct Answer: (A) \( \frac{1}{\pi} \)
View Solution




Step 1: Understanding the Concept:

We need to find the expected value of the area of a triangle whose geometry depends on a random variable \(\Theta\). First, we express the area of the triangle as a function of \(\Theta\), and then we compute its expected value using the distribution of \(\Theta\).




Step 2: Finding the Area of the Triangle as a Function of \(\Theta\):

The vertices of the triangle ABD are \(A=(0,0)\), \(B=(1,0)\), and \(D=(\cos\Theta, \sin\Theta)\).
We can use the formula for the area of a triangle with one vertex at the origin: Area = \(\frac{1}{2} |x_1 y_2 - x_2 y_1|\).
Here, let \(A\) be the origin, \(B=(x_1, y_1)=(1,0)\), and \(D=(x_2, y_2)=(\cos\Theta, \sin\Theta)\). \[ Area(\Theta) = \frac{1}{2} |1 \cdot \sin\Theta - \cos\Theta \cdot 0| = \frac{1}{2} |\sin\Theta| \]
Since \(\Theta \sim U(0, \pi)\), the angle \(\Theta\) is in the first or second quadrant, where \(\sin\Theta \ge 0\).
Therefore, \(|\sin\Theta| = \sin\Theta\).
The area of the triangle is a random variable given by \(Area = \frac{1}{2} \sin\Theta\).

Alternatively, using the formula Area = \(\frac{1}{2}ab\sin C\). The sides adjacent to angle A are AB and AD. The length of AB is the distance from (0,0) to (1,0), which is 1. The length of AD is the distance from (0,0) to \((\cos\Theta, \sin\Theta)\), which is \(\sqrt{\cos^2\Theta+\sin^2\Theta} = 1\) (since D is on the unit circle). The angle between these sides is \(\angle DAB = \Theta\). \[ Area = \frac{1}{2} |AB| \cdot |AD| \sin(\angle DAB) = \frac{1}{2} (1)(1) \sin\Theta = \frac{1}{2} \sin\Theta \]



Step 3: Calculating the Expected Area:

The expected value of a function \(g(\Theta)\) of a continuous random variable \(\Theta\) with PDF \(f(\theta)\) is \(E[g(\Theta)] = \int g(\theta) f(\theta) \, d\theta\).
Here, \(\Theta \sim U(0, \pi)\), so its PDF is \(f(\theta) = \frac{1}{\pi - 0} = \frac{1}{\pi}\) for \(0 < \theta < \pi\), and 0 otherwise.
The function is \(g(\Theta) = Area = \frac{1}{2}\sin\Theta\).
The expected area is: \[ E[Area] = \int_{0}^{\pi} \left(\frac{1}{2}\sin\theta\right) \frac{1}{\pi} \, d\theta = \frac{1}{2\pi} \int_{0}^{\pi} \sin\theta \, d\theta \] \[ = \frac{1}{2\pi} [-\cos\theta]_{0}^{\pi} = \frac{1}{2\pi} (-\cos(\pi) - (-\cos(0))) \] \[ = \frac{1}{2\pi} (-(-1) - (-1)) = \frac{1}{2\pi} (1+1) = \frac{2}{2\pi} = \frac{1}{\pi} \]



Step 4: Final Answer:

The expected area of the triangle ABD is \(\frac{1}{\pi}\).
Quick Tip: When dealing with geometric probability, the first step is always to express the quantity of interest (like area, length, etc.) as a function of the random variable(s). Then, calculate the expectation of this function using the standard integral definition. Visualizing the setup can be very helpful.


Question 19:

Suppose \(Y \sim U(0,1)\) and the conditional distribution of \(X\) given \(Y=y\) is Bin(6, y), for \(0 < y < 1\). Then the probability that \((X+1)\) is an even number is

  • (A) \( \frac{3}{7} \)
  • (B) \( \frac{1}{2} \)
  • (C) \( \frac{4}{7} \)
  • (D) \( \frac{5}{14} \)
Correct Answer: (A) \( \frac{3}{7} \)
View Solution




Step 1: Understanding the Concept:

We need to find the unconditional probability of an event involving \(X\). We can do this using the law of total probability, by first finding the conditional probability of the event given \(Y=y\), and then integrating (or "averaging") this conditional probability over all possible values of \(y\).
The event is \(\{(X+1) is an even number\}\). This is equivalent to the event \(\{X is an odd number\}\).
So, we need to calculate \(P(X is odd)\).




Step 2: Finding the Conditional Probability:

Let's find \(P(X is odd | Y=y)\).
Given \(Y=y\), we have \(X \sim Bin(n=6, p=y)\). The possible values of \(X\) are \(\{0, 1, 2, 3, 4, 5, 6\}\).
We want the probability that \(X\) is odd, i.e., \(X \in \{1, 3, 5\}\). \[ P(X is odd | Y=y) = P(X=1|Y=y) + P(X=3|Y=y) + P(X=5|Y=y) \]
The PMF for Bin(n, p) is \(P(X=k) = \binom{n}{k} p^k (1-p)^{n-k}\). Here \(n=6\) and \(p=y\).
A useful identity for binomial probabilities is: \[ \sum_{k is odd} \binom{n}{k} p^k (1-p)^{n-k} = \frac{1}{2} \left[ 1 - (1-2p)^n \right] \]
Let's verify this. Consider the expansion of \((q+p)^n\) and \((q-p)^n\), where \(q=1-p\). \((q+p)^n = \sum \binom{n}{k} p^k q^{n-k}\) \((q-p)^n = \sum \binom{n}{k} (-p)^k q^{n-k} = \sum \binom{n}{k} (-1)^k p^k q^{n-k}\)
Subtracting the second from the first: \((q+p)^n - (q-p)^n = 2 \sum_{k is odd} \binom{n}{k} p^k q^{n-k}\)
Since \(q+p=1\) and \(q-p=1-p-p=1-2p\), we have: \[ 1^n - (1-2p)^n = 2 P(X is odd) \] \[ P(X is odd) = \frac{1 - (1-2p)^n}{2} \]
In our case, \(n=6\) and \(p=y\). So, \[ P(X is odd | Y=y) = \frac{1 - (1-2y)^6}{2} \]



Step 3: Calculating the Unconditional Probability:

The function \((1-2y)^6\) is symmetric around \(y=1/2\) on \([0,1]\).
\(\int_0^1 (1-2y)^6 dy = \int_0^{1/2} (1-2y)^6 dy + \int_{1/2}^1 (1-2y)^6 dy\).
\(\int_0^1 (1-2y)^6 dy = [ \frac{(1-2y)^7}{7 \cdot (-2)} ]_0^1 = -\frac{1}{14} [ (1-2)^7 - (1-0)^7 ] = -\frac{1}{14}[(-1)^7 - 1^7] = -\frac{1}{14}[-1-1] = -\frac{-2}{14} = \frac{1}{7}\). Correct.

So, putting it all together: \[ P(X is odd) = \frac{1}{2} \left[ 1 - \frac{1}{7} \right] = \frac{1}{2} \left[ \frac{6}{7} \right] = \frac{3}{7} \]
So the answer is 3/7. Quick Tip: This is a classic example of a Beta-Binomial model. When the prior distribution for the success probability \(p\) of a binomial distribution is a Beta(\(\alpha, \beta\)) distribution, the marginal (unconditional) distribution of the number of successes is a Beta-Binomial distribution. A special case is when the prior is U(0,1), which is Beta(1,1). This leads to a simple discrete uniform distribution for the marginal of \(X\).


Question 20:

Let \(X\) be an Exp(\(\lambda\)) random variable. Suppose \(Y = \min\{X, 2\}\). Let \(F_X\) and \(F_Y\) denote the distribution functions of \(X\) and \(Y\) respectively. Then which of the following statements is true?

  • (A) \(F_Y\) is a continuous function
  • (B) \(F_Y(y)\) is discontinuous at \(y=2\)
  • (C) \(F_Y(t) \le F_X(t)\) for all \(t \in \mathbb{R}\)
  • (D) \(E(Y) > E(X)\)
Correct Answer: (B) \(F_Y(y)\) is discontinuous at \(y=2\)
View Solution




Step 1: Understanding the Concept:

We are dealing with a transformation of a random variable. Specifically, \(Y\) is a "censored" or "capped" version of \(X\). When a continuous random variable is capped, it often introduces a discrete component into the distribution of the new variable. We need to find the cumulative distribution function (CDF) of \(Y\) and analyze its properties.

The CDF of \(X \sim Exp(\lambda)\) is \(F_X(x) = 1 - e^{-\lambda x}\) for \(x \ge 0\). The mean is \(E[X] = 1/\lambda\).




Step 2: Finding the CDF of Y:

Let's find \(F_Y(y) = P(Y \le y)\). The support of \(Y\) is \((0, 2]\).
For \(y < 0\), \(F_Y(y) = 0\).
For \(0 \le y < 2\): \[ F_Y(y) = P(Y \le y) = P(\min\{X, 2\} \le y) \]
Since \(y < 2\), the condition \(\min\{X, 2\} \le y\) is equivalent to \(X \le y\). \[ F_Y(y) = P(X \le y) = F_X(y) = 1 - e^{-\lambda y} \]
At \(y=2\):
The variable \(Y\) can take the value 2. This happens whenever \(X \ge 2\). \[ P(Y=2) = P(\min\{X, 2\} = 2) = P(X \ge 2) = 1 - F_X(2) = 1 - (1 - e^{-2\lambda}) = e^{-2\lambda} \]
Since there is a positive probability mass at the point \(y=2\), the random variable \(Y\) is not purely continuous. It is a mixed random variable.
For \(y \ge 2\): \[ F_Y(y) = P(Y \le y) = 1 \]
Because the maximum value \(Y\) can take is 2, for any \(y \ge 2\), the event \(Y \le y\) is certain.
So, the CDF of \(Y\) is:

Step 3: Analyzing the Statements:

(A) \(F_Y\) is a continuous function.
Let's check for continuity at \(y=2\).
The limit from the left is \(\lim_{y \to 2^-} F_Y(y) = \lim_{y \to 2^-} (1 - e^{-\lambda y}) = 1 - e^{-2\lambda}\).
The value at \(y=2\) is \(F_Y(2) = P(Y \le 2) = 1\).
Since \(1 - e^{-2\lambda} \ne 1\) (as \(\lambda > 0\)), the function is not continuous at \(y=2\). So, (A) is false.

(B) \(F_Y(y)\) is discontinuous at \(y=2\).
As shown above, the left-hand limit at \(y=2\) is \(1 - e^{-2\lambda}\) while the value of the function is \(F_Y(2)=1\). The limit does not equal the function value, so the CDF is discontinuous at \(y=2\). This statement is true.

(C) \(F_Y(t) \le F_X(t)\) for all \(t \in \mathbb{R}\).
Let's check this. \(Y = \min\{X, 2\}\) implies that \(Y \le X\) always.
If one random variable is always less than or equal to another (\(Y \le X\)), then their CDFs must satisfy \(F_Y(t) \ge F_X(t)\) for all \(t\). This is because the event \(\{X \le t\}\) is a subset of the event \(\{Y \le t\}\).
So, \(P(X \le t) \le P(Y \le t)\).
Thus, \(F_X(t) \le F_Y(t)\). The statement in the option is the reverse inequality. So, (C) is false.

(D) \(E(Y) > E(X)\).
Since \(Y = \min\{X, 2\}\), we always have \(Y \le X\).
This implies that \(E[Y] \le E[X]\). The inequality can be strict if \(P(X>2) > 0\), which is true for any exponential distribution.
Therefore, \(E[Y] < E[X]\). The statement says the opposite. So, (D) is false.



Step 4: Final Answer:

The only true statement is (B). Quick Tip: When a continuous random variable \(X\) is transformed by \(Y = \min(X, c)\) or \(Y = \max(X, c)\) for some constant \(c\), the resulting variable \(Y\) is a mixed random variable. It has a continuous part inherited from \(X\) and a discrete part (a point mass) at \(y=c\). This point mass will cause a jump discontinuity in the CDF at \(y=c\).


Question 21:


Let \(X_1, X_2, \dots, X_n\) be i.i.d. \(N(0, \sigma^2)\) random variables. Suppose \(c\) is such that \[ E\left[c \sqrt{\sum_{i=1}^n X_i^2}\right] = \sigma. \]
Then the value of \(c\) is

  • (A) \( \sqrt{\frac{9\pi}{128}} \)
  • (B) \( \sqrt{\frac{9\pi}{64}} \)
  • (C) \( \sqrt{\frac{9}{128\pi}} \)
  • (D) \( \sqrt{\frac{3}{64\pi}} \)
Correct Answer: (A) \( \sqrt{\frac{9\pi}{128}} \)
View Solution




Step 1: Understanding the Concept:

The problem asks to find a constant \(c\) that satisfies a given expectation equation involving the square root of the sum of squares of i.i.d. normal random variables. This sum is related to the chi-squared distribution, and its square root is related to the chi distribution.


Step 2: Key Formula or Approach:

1. Standardize the random variables: If \(X_i \sim N(0, \sigma^2)\), then \(Z_i = X_i/\sigma \sim N(0,1)\).

2. The sum of squares of \(n\) standard normal variables follows a chi-squared distribution with \(n\) degrees of freedom: \( \sum_{i=1}^n Z_i^2 \sim \chi^2_n \).

3. The square root of a \(\chi^2_n\) variable is a chi-distributed variable, \(\chi_n\).

4. The expectation of a chi-distributed random variable with \(n\) degrees of freedom is given by: \[ E[\chi_n] = \sqrt{2} \frac{\Gamma((n+1)/2)}{\Gamma(n/2)} \]
where \(\Gamma(z)\) is the gamma function.


Step 3: Detailed Explanation:

The given equation is \( E\left[c \sqrt{\sum_{i=1}^n X_i^2}\right] = \sigma \).

We can rewrite the term inside the expectation using the standardized variables \(Z_i = X_i/\sigma\): \[ \sum_{i=1}^n X_i^2 = \sum_{i=1}^n (\sigma Z_i)^2 = \sigma^2 \sum_{i=1}^n Z_i^2 \]
Substituting this into the equation: \[ E\left[c \sqrt{\sigma^2 \sum_{i=1}^n Z_i^2}\right] = \sigma \] \[ E\left[c \sigma \sqrt{\sum_{i=1}^n Z_i^2}\right] = \sigma \]
Since \(c\) and \(\sigma\) are constants, we can take them out of the expectation: \[ c \sigma E\left[\sqrt{\sum_{i=1}^n Z_i^2}\right] = \sigma \]
Assuming \(\sigma > 0\), we can divide both sides by \(\sigma\): \[ c E\left[\sqrt{\chi^2_n}\right] = 1 \] \[ c E[\chi_n] = 1 \implies c = \frac{1}{E[\chi_n]} \]
Using the formula for the expectation of a chi-distributed variable: \[ c = \frac{1}{\sqrt{2} \frac{\Gamma((n+1)/2)}{\Gamma(n/2)}} = \frac{1}{\sqrt{2}} \frac{\Gamma(n/2)}{\Gamma((n+1)/2)} \]
The value of \(c\) depends on \(n\). The question text has a typo ("X,"), which should be \(X_n\). Since the options are constants, \(n\) must be a specific integer. We can deduce \(n\) by testing which value gives one of the options. Let's test small values of \(n\).

Try \(n=5\): \[ c = \frac{1}{\sqrt{2}} \frac{\Gamma(5/2)}{\Gamma(6/2)} = \frac{1}{\sqrt{2}} \frac{\Gamma(2.5)}{\Gamma(3)} \]
We know \(\Gamma(3) = 2! = 2\) and \(\Gamma(z+1) = z\Gamma(z)\). \[ \Gamma(2.5) = \Gamma(3/2 + 1) = \frac{3}{2} \Gamma(3/2) = \frac{3}{2} \cdot \frac{1}{2} \Gamma(1/2) = \frac{3}{4}\sqrt{\pi} \]
Substituting these values: \[ c = \frac{1}{\sqrt{2}} \frac{(3/4)\sqrt{\pi}}{2} = \frac{3\sqrt{\pi}}{8\sqrt{2}} \]
To match this with the options, let's simplify the square root in option (A): \[ \sqrt{\frac{9\pi}{128}} = \frac{\sqrt{9\pi}}{\sqrt{128}} = \frac{3\sqrt{\pi}}{\sqrt{64 \cdot 2}} = \frac{3\sqrt{\pi}}{8\sqrt{2}} \]
The calculated value for \(c\) with \(n=5\) matches option (A). Thus, the number of random variables is implicitly \(n=5\).


Step 4: Final Answer:

The constant \(c\) is \(\frac{3\sqrt{\pi}}{8\sqrt{2}}\), which is equal to \(\sqrt{\frac{9\pi}{128}}\).
Quick Tip: In problems where a parameter like \(n\) seems to be missing but the answers are numerical, it's often implicitly defined. Test small integer values for \(n\) to see if one matches the given options. The properties of the Gamma function, especially \(\Gamma(n)=(n-1)!\) for integer \(n\) and \(\Gamma(x+1)=x\Gamma(x)\), are essential here.


Question 22:


Let \(X_1, X_2, \dots, X_n\) be a random sample from a continuous distribution with probability density function \[ f(x; \theta) = \frac{1}{2\theta} e^{-|x|/\theta}, \quad x \in \mathbb{R}, \]
where \(\theta > 0\) is an unknown parameter. The critical region for the uniformly most powerful test for testing the null hypothesis \(H_0 : \theta = 2\) against the alternative hypothesis \(H_1 : \theta > 2\) at level \(\alpha\), where \(0 < \alpha < 1\), is

  • (A) \( \left\{ (x_1, \dots, x_n) \in \mathbb{R}^n : 2\sum_{i=1}^n |x_i| < \chi^2_{2n, 1-\alpha} \right\} \)
  • (B) \( \left\{ (x_1, \dots, x_n) \in \mathbb{R}^n : 2\sum_{i=1}^n |x_i| > \chi^2_{2n, \alpha} \right\} \)
  • (C) \( \left\{ (x_1, \dots, x_n) \in \mathbb{R}^n : \sum_{i=1}^n |x_i| < \chi^2_{2n, 1-\alpha} \right\} \)
  • (D) \( \left\{ (x_1, \dots, x_n) \in \mathbb{R}^n : \sum_{i=1}^n |x_i| > \chi^2_{2n, \alpha} \right\} \)
Correct Answer: (D) \( \left\{ (x_1, \dots, x_n) \in \mathbb{R}^n : \sum_{i=1}^n |x_i| > \chi^2_{2n, \alpha} \right\} \)
View Solution




Step 1: Understanding the Concept:

This problem requires finding the Uniformly Most Powerful (UMP) test for a one-sided hypothesis about the parameter \(\theta\) of a Laplace (double exponential) distribution. The Karlin-Rubin theorem is the key tool, which applies to distributions with a Monotone Likelihood Ratio (MLR).


Step 2: Key Formula or Approach:

1. Write down the likelihood function \(L(\theta; \mathbf{x})\).

2. Check if the family of distributions has the MLR property in some statistic \(T(\mathbf{x})\). For \(\theta_2 > \theta_1\), the ratio \(L(\theta_2; \mathbf{x})/L(\theta_1; \mathbf{x})\) should be a non-decreasing function of \(T(\mathbf{x})\).

3. Apply the Karlin-Rubin theorem: For testing \(H_0: \theta = \theta_0\) vs \(H_1: \theta > \theta_0\), the UMP test rejects \(H_0\) for large values of \(T(\mathbf{x})\), i.e., \(T(\mathbf{x}) > k\).

4. Determine the distribution of the test statistic \(T(\mathbf{x})\) under \(H_0\) to find the critical value \(k\) for a given significance level \(\alpha\).


Step 3: Detailed Explanation:

The likelihood function for the random sample \(\mathbf{x} = (x_1, \dots, x_n)\) is: \[ L(\theta; \mathbf{x}) = \prod_{i=1}^n f(x_i; \theta) = \prod_{i=1}^n \frac{1}{2\theta} e^{-|x_i|/\theta} = \left(\frac{1}{2\theta}\right)^n \exp\left(-\frac{1}{\theta} \sum_{i=1}^n |x_i|\right) \]
To check for MLR, let \(\theta_2 > \theta_1\). The likelihood ratio is: \[ \frac{L(\theta_2; \mathbf{x})}{L(\theta_1; \mathbf{x})} = \frac{(1/(2\theta_2))^n \exp(-\frac{1}{\theta_2}\sum|x_i|)}{(1/(2\theta_1))^n \exp(-\frac{1}{\theta_1}\sum|x_i|)} = \left(\frac{\theta_1}{\theta_2}\right)^n \exp\left[\left(\frac{1}{\theta_1} - \frac{1}{\theta_2}\right)\sum_{i=1}^n|x_i|\right] \]
Since \(\theta_2 > \theta_1 > 0\), the term \((\frac{1}{\theta_1} - \frac{1}{\theta_2})\) is positive. Therefore, the likelihood ratio is an increasing function of \(T(\mathbf{x}) = \sum_{i=1}^n |x_i|\). The family of distributions has MLR in \(T(\mathbf{x})\).

By the Karlin-Rubin theorem, the UMP test for \(H_0: \theta = 2\) vs \(H_1: \theta > 2\) rejects \(H_0\) for large values of \(T(\mathbf{x})\). The critical region is of the form \(\left\{ \mathbf{x} : \sum_{i=1}^n |x_i| > k \right\}\).

To find \(k\), we need the distribution of \(T(\mathbf{x})\) under \(H_0\). Let \(Y_i = |X_i|\). The PDF of \(Y_i\) for \(y_i>0\) is \(f_{Y_i}(y_i) = 2 \cdot f_X(y_i) = 2 \frac{1}{2\theta} e^{-y_i/\theta} = \frac{1}{\theta} e^{-y_i/\theta}\), which is an exponential distribution with mean \(\theta\).

The sum of \(n\) i.i.d. Exponential(\(rate=\lambda\)) variables follows a Gamma(\(n, rate=\lambda\)) distribution. Here the rate is \(\lambda = 1/\theta\), so \(\sum |X_i| \sim Gamma(n, scale=\theta)\).

We know that if \(W \sim Gamma(shape=\alpha, scale=\beta)\), then \(2W/\beta \sim \chi^2_{2\alpha}\).

Here, \(T = \sum |X_i|\), \(\alpha=n\), \(\beta=\theta\). So, \(2T/\theta = \frac{2\sum |X_i|}{\theta} \sim \chi^2_{2n}\).

Under \(H_0\), we have \(\theta=2\). The test statistic becomes \(\frac{2\sum |X_i|}{2} = \sum |X_i|\). Thus, under \(H_0\), \(\sum_{i=1}^n |X_i| \sim \chi^2_{2n}\).

The critical value \(k\) is determined by the level \(\alpha\): \[ P_{H_0}\left(\sum_{i=1}^n |X_i| > k\right) = \alpha \]
Since \(\sum |X_i|\) follows a \(\chi^2_{2n}\) distribution under \(H_0\), \(k\) is the upper \(\alpha\)-quantile of this distribution, denoted as \(\chi^2_{2n, \alpha}\).


Step 4: Final Answer:

The critical region for the UMP test is \(\left\{ (x_1, \dots, x_n) \in \mathbb{R}^n : \sum_{i=1}^n |x_i| > \chi^2_{2n, \alpha} \right\}\), which matches option (D).
Quick Tip: For one-sided hypothesis tests involving a single parameter, the Karlin-Rubin theorem is your go-to tool. The main steps are to check for the Monotone Likelihood Ratio (MLR) property and then identify the distribution of the resulting test statistic under the null hypothesis to find the critical value.


Question 23:


Let \((X, Y)\) have the \(N_2(0,0,1,1,0.25)\) distribution. Then the correlation coefficient between \(e^X\) and \(e^{2Y}\) is

  • (A) \( \frac{e^3 - e^{5/2}}{(e^5(e-1)(e^4-1))^{1/2}} \)
  • (B) \( \frac{e^3 - e^{5/2}}{(e^4(e-1)(e^8-1))^{1/2}} \)
  • (C) \( \frac{e^2 - e^{5/2}}{(e^5(e-1)(e^4-1))^{1/2}} \)
  • (D) \( \frac{e^{3/2} - e^{5/2}}{(e^5(e^2-1)(e^4-1))^{1/2}} \)
Correct Answer: (A) \( \frac{e^3 - e^{5/2}}{(e^5(e-1)(e^4-1))^{1/2}} \)
View Solution




Step 1: Understanding the Concept:

The problem asks for the correlation coefficient between two functions of random variables that follow a bivariate normal distribution. This requires calculating their covariance and variances. The key tool is the moment generating function (MGF) of normal and bivariate normal distributions.


Step 2: Key Formula or Approach:

1. Correlation coefficient: \(Corr(U,V) = \frac{Cov(U,V)}{\sqrt{Var(U)Var(V)}}\).

2. Covariance: \(Cov(U,V) = E[UV] - E[U]E[V]\).

3. Variance: \(Var(U) = E[U^2] - (E[U])^2\).

4. MGF of a normal variable \(Z \sim N(\mu, \sigma^2)\): \(M_Z(t) = E[e^{tZ}] = e^{t\mu + t^2\sigma^2/2}\).

5. If \((X, Y)\) are bivariate normal, then any linear combination \(W = aX + bY\) is also normal.

\(E[W] = aE[X] + bE[Y]\).

\(Var(W) = a^2Var(X) + b^2Var(Y) + 2abCov(X,Y)\).


Step 3: Detailed Explanation:

Given \((X, Y) \sim N_2(0,0,1,1,0.25)\), we have \(\mu_X=0, \mu_Y=0, \sigma_X^2=1, \sigma_Y^2=1, \rho=0.25\).
Thus, \(X \sim N(0,1)\) and \(Y \sim N(0,1)\). Let \(U=e^X\) and \(V=e^{2Y}\).

Calculate moments for U = \(e^X\): \(E[U] = E[e^X] = M_X(1) = e^{1 \cdot 0 + 1^2 \cdot 1 / 2} = e^{1/2}\).
\(E[U^2] = E[e^{2X}] = M_X(2) = e^{2 \cdot 0 + 2^2 \cdot 1 / 2} = e^2\).
\(Var(U) = E[U^2] - (E[U])^2 = e^2 - (e^{1/2})^2 = e^2 - e = e(e-1)\).


Calculate moments for V = \(e^{2Y}\): \(E[V] = E[e^{2Y}] = M_Y(2) = e^{2 \cdot 0 + 2^2 \cdot 1 / 2} = e^2\).
\(E[V^2] = E[(e^{2Y})^2] = E[e^{4Y}] = M_Y(4) = e^{4 \cdot 0 + 4^2 \cdot 1 / 2} = e^8\).
\(Var(V) = E[V^2] - (E[V])^2 = e^8 - (e^2)^2 = e^8 - e^4 = e^4(e^4-1)\).


Calculate \(E[UV]\): \(E[UV] = E[e^X e^{2Y}] = E[e^{X+2Y}]\).
Let \(W = X+2Y\). Since \((X,Y)\) is bivariate normal, \(W\) is normal. \(E[W] = E[X] + 2E[Y] = 0 + 2(0) = 0\).
\(Cov(X,Y) = \rho \sigma_X \sigma_Y = 0.25 \cdot 1 \cdot 1 = 0.25\).
\(Var(W) = Var(X) + 4Var(Y) + 2 \cdot 1 \cdot 2 \cdot Cov(X,Y) = 1 + 4(1) + 4(0.25) = 1 + 4 + 1 = 6\).

So, \(W \sim N(0, 6)\). \(E[e^{X+2Y}] = E[e^W] = M_W(1) = e^{1 \cdot 0 + 1^2 \cdot 6 / 2} = e^3\).


Calculate Covariance and Correlation: \(Cov(U,V) = E[UV] - E[U]E[V] = e^3 - (e^{1/2})(e^2) = e^3 - e^{5/2}\).
\(Corr(U,V) = \frac{e^3 - e^{5/2}}{\sqrt{Var(U)Var(V)}} = \frac{e^3 - e^{5/2}}{\sqrt{e(e-1) \cdot e^4(e^4-1)}} = \frac{e^3 - e^{5/2}}{\sqrt{e^5(e-1)(e^4-1)}}\).


Step 4: Final Answer:

The correlation coefficient is \(\frac{e^3 - e^{5/2}}{(e^5(e-1)(e^4-1))^{1/2}}\), which corresponds to option (A).
Quick Tip: When dealing with expectations of exponential functions of normal random variables, remember that you are essentially calculating values of the Moment Generating Function (MGF). For a linear combination like \(aX+bY\), first find the distribution of the combination (which will be normal) and then use its MGF.


Question 24:


Let \(\{X_k\}_{k \geq 1}\) be a sequence of i.i.d. U(-1,1) random variables. Suppose \[ Y_n = \sqrt{3n} \frac{\sum_{i=1}^n X_i}{\sum_{i=1}^n X_i^4}. \]
Then \(\{Y_n\}_{n \geq 1}\) converges in distribution as \(n \to \infty\) to a

  • (A) \(N(0,1)\) random variable
  • (B) random variable degenerate at 0
  • (C) \(N(0,25)\) random variable
  • (D) \(N(0,0.04)\) random variable
Correct Answer: (C) \(N(0,25)\) random variable
View Solution




Step 1: Understanding the Concept:

The problem asks for the limiting distribution of a sequence of random variables \(Y_n\). The structure of \(Y_n\) involves sums of i.i.d. random variables in both the numerator and denominator, suggesting the use of the Central Limit Theorem (CLT), the Law of Large Numbers (LLN), and Slutsky's Theorem.


Step 2: Key Formula or Approach:

1. Calculate the necessary moments of the \(X_k\) variables: \(E[X_k]\), \(Var(X_k)\), and \(E[X_k^4]\).

2. Rewrite \(Y_n\) to isolate terms that have known convergence properties.
\[ Y_n = \sqrt{3} \frac{\sqrt{n}(\frac{1}{n}\sum_{i=1}^n X_i)}{\frac{1}{n}\sum_{i=1}^n X_i^4} \]
3. Apply the Central Limit Theorem to the numerator term.

4. Apply the Weak Law of Large Numbers (WLLN) to the denominator term.

5. Use Slutsky's Theorem to find the limiting distribution of the ratio. Slutsky's Theorem states that if \(A_n \xrightarrow{d} A\) and \(B_n \xrightarrow{p} b\) (a constant), then \(A_n/B_n \xrightarrow{d} A/b\).


Step 3: Detailed Explanation:

First, we find the moments for \(X_k \sim U(-1,1)\). The PDF is \(f(x) = 1/2\) for \(x \in [-1,1]\). \(E[X_k] = \int_{-1}^1 x \cdot \frac{1}{2} dx = 0\).
\(E[X_k^2] = \int_{-1}^1 x^2 \cdot \frac{1}{2} dx = \frac{1}{2}\left[\frac{x^3}{3}\right]_{-1}^1 = \frac{1}{3}\).
\(Var(X_k) = E[X_k^2] - (E[X_k])^2 = 1/3\).
\(E[X_k^4] = \int_{-1}^1 x^4 \cdot \frac{1}{2} dx = \frac{1}{2}\left[\frac{x^5}{5}\right]_{-1}^1 = \frac{1}{5}\).


Now let's analyze \(Y_n\). We can write it as: \[ Y_n = \sqrt{3} \cdot \frac{\sqrt{n} \bar{X}_n}{\frac{1}{n}\sum_{i=1}^n X_i^4} \]
Let's analyze the numerator and denominator separately.

Numerator: Let \(A_n = \sqrt{n} \bar{X}_n = \frac{1}{\sqrt{n}}\sum_{i=1}^n X_i\). By the Central Limit Theorem: \[ \frac{\sum X_i - n\mu}{\sqrt{n\sigma^2}} = \frac{\sqrt{n}(\bar{X}_n - \mu)}{\sigma} \xrightarrow{d} N(0,1) \]
Here \(\mu=0\) and \(\sigma^2=1/3\). \[ \frac{\sqrt{n}\bar{X}_n}{1/\sqrt{3}} \xrightarrow{d} N(0,1) \implies \sqrt{3n}\bar{X}_n \xrightarrow{d} N(0,1) \]
So, \(\sqrt{n}\bar{X}_n \xrightarrow{d} N(0, 1/3)\). Let's call the numerator \(N_n = \sqrt{n}\bar{X}_n\).

Denominator: Let \(D_n = \frac{1}{n}\sum_{i=1}^n X_i^4\). By the Weak Law of Large Numbers: \[ D_n \xrightarrow{p} E[X_i^4] = 1/5 \]
Now we apply Slutsky's Theorem to \(Y_n = \sqrt{3} \frac{N_n}{D_n}\).
Since \(N_n \xrightarrow{d} N(0, 1/3)\) and \(D_n \xrightarrow{p} 1/5\), we have: \[ \frac{N_n}{D_n} \xrightarrow{d} \frac{N(0, 1/3)}{1/5} = 5 \cdot N(0, 1/3) \]
Using the property that \(c \cdot N(\mu, \sigma^2) \sim N(c\mu, c^2\sigma^2)\): \[ 5 \cdot N(0, 1/3) \sim N(5 \cdot 0, 5^2 \cdot (1/3)) = N(0, 25/3) \]
Finally, we find the distribution of \(Y_n\): \[ Y_n = \sqrt{3} \cdot \frac{N_n}{D_n} \xrightarrow{d} \sqrt{3} \cdot N(0, 25/3) \] \[ \sqrt{3} \cdot N(0, 25/3) \sim N(\sqrt{3} \cdot 0, (\sqrt{3})^2 \cdot (25/3)) = N(0, 3 \cdot 25/3) = N(0, 25) \]

Step 4: Final Answer:

The sequence \(\{Y_n\}\) converges in distribution to a \(N(0,25)\) random variable. This corresponds to option (C).
Quick Tip: When faced with a complex ratio of sums of random variables, the combination of CLT for the numerator (if properly scaled by \(\sqrt{n}\)) and LLN for the denominator (if scaled by \(n\)) is a powerful strategy. Slutsky's Theorem is the final piece that allows you to combine these results.


Question 25:


If \((X, Y) \sim N_2(0,0,1,1,0.5)\), then the value of \(E[e^{-XY}]\) is

  • (A) \( \frac{2}{\sqrt{5}} \)
  • (B) \( \frac{2}{\sqrt{3}} \)
  • (C) \( \frac{1}{\sqrt{2}} \)
  • (D) \( \frac{1}{2} \)
Correct Answer: (A) \( \frac{2}{\sqrt{5}} \)
View Solution




Step 1: Understanding the Concept:

This problem asks for the expectation of a non-linear function of two jointly normal random variables. A direct integration would be complicated. A more effective method is to use the law of iterated expectations (tower property), by first conditioning on one of the variables.


Step 2: Key Formula or Approach:

1. Law of Iterated Expectations: \(E[g(X,Y)] = E_X[E_{Y|X}[g(X,Y)|X=x]]\).

2. Conditional Distribution for Bivariate Normal: If \((X,Y) \sim N(\mu_X, \mu_Y, \sigma_X^2, \sigma_Y^2, \rho)\), then the conditional distribution of \(Y\) given \(X=x\) is also normal:
\[ Y|X=x \sim N\left(\mu_Y + \rho\frac{\sigma_Y}{\sigma_X}(x-\mu_X), \sigma_Y^2(1-\rho^2)\right) \]
3. MGF of a Normal Variable \(Z \sim N(\mu, \sigma^2)\): \(M_Z(t) = E[e^{tZ}] = e^{t\mu + t^2\sigma^2/2}\).

4. Integral of a Gaussian kernel: \(\int_{-\infty}^{\infty} e^{-ax^2} dx = \sqrt{\frac{\pi}{a}}\).


Step 3: Detailed Explanation:

We want to calculate \(E[e^{-XY}]\). Using iterated expectations: \(E[e^{-XY}] = E_X[E[e^{-XY}|X=x]]\).
First, find the conditional distribution of \(Y|X=x\).
Given parameters: \(\mu_X=0, \mu_Y=0, \sigma_X^2=1, \sigma_Y^2=1, \rho=0.5\). \[ Y|X=x \sim N\left(0 + 0.5\frac{1}{1}(x-0), 1(1-0.5^2)\right) = N(0.5x, 0.75) \]
The inner expectation is \(E[e^{-xY}|X=x]\). This is the MGF of the conditional distribution \(Y|X=x\) evaluated at \(t = -x\).
The MGF for \(N(\mu, \sigma^2)\) is \(e^{\mu t + \sigma^2 t^2/2}\). Here, \(\mu = 0.5x\) and \(\sigma^2 = 0.75\). \[ E[e^{-xY}|X=x] = \exp\left((0.5x)(-x) + \frac{0.75(-x)^2}{2}\right) = \exp\left(-0.5x^2 + \frac{0.75}{2}x^2\right) \] \[ = \exp\left(-\frac{1}{2}x^2 + \frac{3/4}{2}x^2\right) = \exp\left(-\frac{1}{2}x^2 + \frac{3}{8}x^2\right) = \exp\left(\frac{-4+3}{8}x^2\right) = e^{-x^2/8} \]
Now, we take the expectation of this result with respect to \(X\), where \(X \sim N(0,1)\). The PDF of \(X\) is \(f_X(x) = \frac{1}{\sqrt{2\pi}}e^{-x^2/2}\). \[ E_X[e^{-X^2/8}] = \int_{-\infty}^{\infty} e^{-x^2/8} f_X(x) dx = \int_{-\infty}^{\infty} e^{-x^2/8} \frac{1}{\sqrt{2\pi}}e^{-x^2/2} dx \] \[ = \frac{1}{\sqrt{2\pi}} \int_{-\infty}^{\infty} \exp\left(-\frac{x^2}{8} - \frac{x^2}{2}\right) dx = \frac{1}{\sqrt{2\pi}} \int_{-\infty}^{\infty} \exp\left(-\frac{x^2 + 4x^2}{8}\right) dx \] \[ = \frac{1}{\sqrt{2\pi}} \int_{-\infty}^{\infty} \exp\left(-\frac{5}{8}x^2\right) dx \]
This integral is of the form \(\int_{-\infty}^\infty e^{-ax^2}dx = \sqrt{\pi/a}\), with \(a = 5/8\). \[ E[e^{-XY}] = \frac{1}{\sqrt{2\pi}} \sqrt{\frac{\pi}{5/8}} = \frac{1}{\sqrt{2\pi}} \sqrt{\frac{8\pi}{5}} = \frac{1}{\sqrt{2\pi}} \frac{\sqrt{8}\sqrt{\pi}}{\sqrt{5}} = \frac{2\sqrt{2}}{\sqrt{2}\sqrt{5}} = \frac{2}{\sqrt{5}} \]

Step 4: Final Answer:

The value of \(E[e^{-XY}]\) is \(\frac{2}{\sqrt{5}}\), which is option (A).
Quick Tip: The law of iterated expectations is an extremely powerful tool for complicated expectations, especially with normal distributions. Conditioning on one variable simplifies the problem to finding an MGF, and the final step often reduces to a standard Gaussian integral.


Question 26:


Let \((X_1, Y_1), (X_2, Y_2), \dots, (X_n, Y_n)\) be a random sample from a \(N_2(0,0,1,1,\rho)\) distribution, where \(\rho\) is an unknown parameter. Which of the following statements is NOT correct?

  • (A) \( \left(\sum_{i=1}^n X_i^2, \sum_{i=1}^n Y_i^2, \sum_{i=1}^n X_i Y_i\right) \) is a sufficient statistic for \(\rho\)
  • (B) \( \left(\sum_{i=1}^n X_i^2, \sum_{i=1}^n Y_i^2, \sum_{i=1}^n X_i Y_i\right) \) is not a minimal sufficient statistic for \(\rho\)
  • (C) \( \sum_{i=1}^n X_i^2 \) is an ancillary statistic
  • (D) \( \left(\sum_{i=1}^n X_i^2, \sum_{i=1}^n Y_i^2\right) \) is an ancillary statistic
Correct Answer: (D) \( \left(\sum_{i=1}^n X_i^2, \sum_{i=1}^n Y_i^2\right) \) is an ancillary statistic
View Solution




Step 1: Understanding the Concept:

This question tests the understanding of key statistical concepts: sufficiency, minimal sufficiency, and ancillarity, in the context of a bivariate normal distribution where only the correlation coefficient \(\rho\) is unknown.


Step 2: Key Formula or Approach:

1. Sufficiency: Use the Fisher-Neyman Factorization Theorem. A statistic \(T(\mathbf{X})\) is sufficient for \(\theta\) if the likelihood function can be factored as \(L(\theta; \mathbf{x}) = g(T(\mathbf{x}); \theta)h(\mathbf{x})\).

2. Minimal Sufficiency: A sufficient statistic is minimal if it is a function of every other sufficient statistic. For exponential families, the natural sufficient statistic is minimal if the parameter space contains an open set. A more general method is to check if \(L(\theta; \mathbf{x}) / L(\theta; \mathbf{y})\) is constant in \(\theta\) iff \(T(\mathbf{x})=T(\mathbf{y})\).

3. Ancillary Statistic: A statistic \(S(\mathbf{X})\) is ancillary for \(\theta\) if its distribution does not depend on \(\theta\).


Step 3: Detailed Explanation:

The PDF for a single observation \((X_i, Y_i)\) from \(N_2(0,0,1,1,\rho)\) is: \[ f(x_i, y_i; \rho) = \frac{1}{2\pi \sqrt{1-\rho^2}} \exp\left( -\frac{1}{2(1-\rho^2)} (x_i^2 - 2\rho x_i y_i + y_i^2) \right) \]
The likelihood for the entire sample is: \[ L(\rho; \mathbf{x, y}) = \left(\frac{1}{2\pi \sqrt{1-\rho^2}}\right)^n \exp\left( -\frac{\sum x_i^2 + \sum y_i^2}{2(1-\rho^2)} + \frac{\rho \sum x_i y_i}{1-\rho^2} \right) \]

(A) Sufficiency:
The likelihood \(L(\rho; \mathbf{x, y})\) depends on the data only through the terms \(\sum X_i^2\), \(\sum Y_i^2\), and \(\sum X_i Y_i\). By the Factorization Theorem, with \(h(\mathbf{x,y})=1\), the statistic \( T = (\sum X_i^2, \sum Y_i^2, \sum X_i Y_i) \) is sufficient for \(\rho\). Statement (A) is correct.


(B) Minimal Sufficiency:
We can write the exponent in the likelihood as: \[ \frac{\rho}{1-\rho^2} \sum x_i y_i - \frac{1}{2(1-\rho^2)} (\sum x_i^2 + \sum y_i^2) \]
This is an exponential family with sufficient statistic \( S = (\sum X_i Y_i, \sum X_i^2 + \sum Y_i^2) \). This statistic \(S\) is minimal sufficient for \(\rho\). The statistic \(T\) from statement (A) is \((\sum X_i^2, \sum Y_i^2, \sum X_i Y_i)\). We can obtain \(S\) from \(T\), but we cannot obtain \(T\) from \(S\) (we can't recover \(\sum X_i^2\) and \(\sum Y_i^2\) individually from their sum). Since \(T\) is not a function of the minimal sufficient statistic \(S\), \(T\) itself cannot be minimal. Therefore, \(T\) is sufficient but not minimal. Statement (B) is correct.


(C) Ancillary Statistic \(\sum X_i^2\):
The marginal distribution of \(X_i\) is \(N(0,1)\), which does not depend on \(\rho\). Therefore, \(X_i^2 \sim \chi^2_1\). The sum of i.i.d. variables, \(\sum_{i=1}^n X_i^2\), follows a \(\chi^2_n\) distribution. Since this distribution does not depend on \(\rho\), \(\sum X_i^2\) is an ancillary statistic. Statement (C) is correct.


(D) Ancillary Statistic \((\sum X_i^2, \sum Y_i^2)\):
For this vector statistic to be ancillary, its joint distribution must not depend on \(\rho\). While the marginal distributions of \(\sum X_i^2\) and \(\sum Y_i^2\) are both \(\chi^2_n\) (independent of \(\rho\)), their joint distribution might depend on \(\rho\). The variables \(X_i\) and \(Y_i\) are correlated with correlation \(\rho\). This means \(X_i^2\) and \(Y_i^2\) are also correlated, and their covariance will depend on \(\rho\). \(Cov(X_i^2, Y_i^2) = E[X_i^2 Y_i^2] - E[X_i^2]E[Y_i^2]\).
For a standard bivariate normal, \(E[X^2 Y^2] = 1 + 2\rho^2\). \(E[X_i^2] = Var(X_i) + (E[X_i])^2 = 1+0=1\). Similarly, \(E[Y_i^2]=1\).
So, \(Cov(X_i^2, Y_i^2) = (1+2\rho^2) - 1 \cdot 1 = 2\rho^2\).
Since the covariance between \(\sum X_i^2\) and \(\sum Y_i^2\) is \(n \cdot 2\rho^2\), which depends on \(\rho\) (unless \(\rho=0\)), their joint distribution depends on \(\rho\). Therefore, \((\sum X_i^2, \sum Y_i^2)\) is NOT an ancillary statistic. Statement (D) is NOT correct.


Step 4: Final Answer:

The question asks for the statement that is NOT correct. Based on the analysis, statement (D) is incorrect.
Quick Tip: To check if a statistic is ancillary, you must verify that its entire probability distribution is free of the parameter. For a vector statistic, this means the joint distribution must be free of the parameter. Even if marginal distributions are parameter-free, the joint distribution may not be if the components are correlated in a parameter-dependent way.


Question 27:

Let \((Y_1, Y_2, Y_3) \in \{0, 1, \dots, n\}^3\) be a discrete random vector having joint probability mass function

where \(0 \le p \le \frac{1}{3}\) is an unknown parameter. Assume the convention \(0^0=1\). The maximum likelihood estimator of \(p\) is denoted by \(\hat{p}\). Which of the following statements is correct?

  • (A) \(E(\hat{p}) > p\)
  • (B) \(\hat{p}\) is an unbiased estimator of \(p\), but not the uniformly minimum variance unbiased estimator of \(p\)
  • (C) \(\hat{p}\) is the uniformly minimum variance unbiased estimator of \(p\)
  • (D) \(E(\hat{p}) < p\)
Correct Answer: (C) \(\hat{p}\) is the uniformly minimum variance unbiased estimator of \(p\)
View Solution




Step 1: Understanding the Concept:

The given joint PMF corresponds to a Multinomial distribution with \(n\) trials and three categories with probabilities \(p_1=p\), \(p_2=2p\), and \(p_3=1-3p\). We need to find the Maximum Likelihood Estimator (MLE) of \(p\), and then determine if it is unbiased and/or the Uniformly Minimum Variance Unbiased Estimator (UMVUE).




Step 2: Finding the MLE of \(p\):

The likelihood function, for a single observation \((y_1, y_2, y_3)\) such that \(y_1+y_2+y_3=n\), is: \[ L(p) = \frac{n!}{y_1! y_2! y_3!} p^{y_1} (2p)^{y_2} (1-3p)^{n-y_1-y_2} \] \[ L(p) \propto p^{y_1} (2p)^{y_2} (1-3p)^{n-y_1-y_2} = 2^{y_2} p^{y_1+y_2} (1-3p)^{n-y_1-y_2} \]
The log-likelihood function is: \[ \ln L(p) = constant + (y_1+y_2)\ln(p) + (n-y_1-y_2)\ln(1-3p) \]
To find the MLE, we differentiate with respect to \(p\) and set the derivative to zero: \[ \frac{d}{dp} \ln L(p) = \frac{y_1+y_2}{p} + \frac{(n-y_1-y_2)(-3)}{1-3p} = 0 \] \[ \frac{y_1+y_2}{p} = \frac{3(n-y_1-y_2)}{1-3p} \] \[ (y_1+y_2)(1-3p) = 3p(n-y_1-y_2) \] \[ y_1+y_2 - 3p(y_1+y_2) = 3np - 3p(y_1+y_2) \] \[ y_1+y_2 = 3np \]
Solving for \(p\), we get the MLE: \[ \hat{p} = \frac{Y_1+Y_2}{3n} \]



Step 3: Checking for Unbiasedness:

We need to calculate the expected value of \(\hat{p}\). In a Multinomial\((n, p_1, p_2, p_3)\) distribution, the marginal distribution of \(Y_i\) is Binomial\((n, p_i)\). Thus, \(E[Y_i] = np_i\). \[ E[Y_1] = n \cdot p_1 = np \] \[ E[Y_2] = n \cdot p_2 = n(2p) = 2np \]
Now, we find the expectation of the estimator: \[ E[\hat{p}] = E\left[\frac{Y_1+Y_2}{3n}\right] = \frac{1}{3n} E[Y_1+Y_2] = \frac{1}{3n} (E[Y_1] + E[Y_2]) \] \[ E[\hat{p}] = \frac{1}{3n} (np + 2np) = \frac{3np}{3n} = p \]
Since \(E[\hat{p}] = p\), the estimator is unbiased. This eliminates options (A) and (D).




Step 4: Checking for UMVUE property:

To check if \(\hat{p}\) is the UMVUE, we can use the Lehmann-Scheffé theorem. We first check if the distribution belongs to the one-parameter exponential family. The PMF can be written as: \[ P(\mathbf{y}; p) = \frac{n!}{y_1! y_2! y_3!} \exp\left( (y_1+y_2)\ln p + y_2\ln 2 + (n-y_1-y_2)\ln(1-3p) \right) \] \[ P(\mathbf{y}; p) = h(\mathbf{y}) c(p) \exp(w(p) T(\mathbf{y})) \]
where \(T(\mathbf{y}) = y_1+y_2\) is a sufficient statistic. This is a regular one-parameter exponential family, for which the sufficient statistic \(T=Y_1+Y_2\) is also complete.
The Lehmann-Scheffé theorem states that if an estimator is a function of a complete sufficient statistic and is unbiased, then it is the UMVUE.
Our estimator \(\hat{p} = \frac{T}{3n}\) is a function of the complete sufficient statistic \(T = Y_1+Y_2\). We have already shown that it is unbiased.
Therefore, \(\hat{p}\) is the UMVUE of \(p\). This means statement (C) is correct and (B) is incorrect.
Quick Tip: For distributions in the exponential family, finding the MLE often leads to a function of the sufficient statistic. If you can show this estimator is unbiased and the family is complete (which is usually true for regular exponential families), the Lehmann-Scheffé theorem is a powerful tool to prove it is the UMVUE.


Question 28:

Let \(X_1, X_2, \dots, X_n\) be i.i.d. \(N(0,1)\) random variables, where \(n > 3\). If \[ \bar{X} = \frac{1}{n} \sum_{i=1}^n X_i \quad and \quad S^2 = \frac{1}{n-1} \sum_{i=1}^n (X_i - \bar{X})^2, \]
then Var\((\frac{\bar{X}}{S})\) is equal to

  • (A) \( \frac{n-3}{n(n-1)} \)
  • (B) \( \frac{n-1}{n(n-3)} \)
  • (C) \( \frac{n-1}{n(n-2)} \)
  • (D) \( \frac{n-2}{n(n-1)} \)
Correct Answer: (B) \( \frac{n-1}{n(n-3)} \)
View Solution




Step 1: Understanding the Concept:

We need to find the variance of the ratio of the sample mean to the sample standard deviation. The key is to use the properties of sampling distributions from a normal population, particularly the independence of \(\bar{X}\) and \(S^2\), and the distributions they follow.




Step 2: Key Formula or Approach:

The variance of a random variable \(Y\) is given by Var(\(Y\)) = \(E[Y^2] - (E[Y])^2\). Let \(Y = \bar{X}/S\).
First, let's find the distributions of \(\bar{X}\) and \(S\).
Since \(X_i \sim i.i.d. N(0,1)\):

\(\bar{X} \sim N(0, 1/n)\).
\((n-1)S^2 = \sum (X_i - \bar{X})^2 \sim \chi^2_{n-1}\) (chi-squared with \(n-1\) degrees of freedom).
By Cochran's theorem, \(\bar{X}\) and \(S^2\) (and thus \(S\)) are independent.




Step 3: Detailed Explanation:

Let's compute the terms for the variance formula.

Finding \(E[\bar{X}/S]\):

Since \(\bar{X}\) and \(S\) are independent, we have \(E[\bar{X}/S] = E[\bar{X}] \cdot E[1/S]\).
The mean of \(\bar{X}\) is \(E[\bar{X}] = 0\).
Therefore, \(E[\bar{X}/S] = 0 \cdot E[1/S] = 0\).

Finding \(E[(\bar{X}/S)^2]\):

Since \(E[\bar{X}/S] = 0\), the variance is simply \(Var(\bar{X}/S) = E[(\bar{X}/S)^2]\).
Using the independence of \(\bar{X}\) and \(S\) again:
\[ E\left[\left(\frac{\bar{X}}{S}\right)^2\right] = E\left[\frac{\bar{X}^2}{S^2}\right] = E[\bar{X}^2] \cdot E\left[\frac{1}{S^2}\right] \]
Now we calculate the two expectations separately.

\(E[\bar{X}^2]\): We know \(Var(\bar{X}) = E[\bar{X}^2] - (E[\bar{X}])^2\).
Since \(E[\bar{X}] = 0\) and \(Var(\bar{X}) = 1/n\), we have \(E[\bar{X}^2] = 1/n\).

\(E[1/S^2]\): Let \(W = (n-1)S^2 \sim \chi^2_{n-1}\). Then \(S^2 = W/(n-1)\), so \(1/S^2 = (n-1)/W\).
We need \(E[1/S^2] = E[(n-1)/W] = (n-1)E[1/W]\).
For a random variable \(W \sim \chi^2_k\), the expectation \(E[W^m]\) is given by \(\frac{2^m \Gamma(k/2+m)}{\Gamma(k/2)}\). For \(m=-1\), we need to find \(E[W^{-1}]\).
A simpler result is that if \(W \sim \chi^2_k\), then \(E[1/W] = \frac{1}{k-2}\) for \(k > 2\).
Here, our degrees of freedom are \(k = n-1\). The condition \(n-1 > 2\) implies \(n>3\), which is given in the problem.
So, \(E[1/W] = \frac{1}{(n-1)-2} = \frac{1}{n-3}\).
Therefore, \(E[1/S^2] = (n-1) \cdot E[1/W] = \frac{n-1}{n-3}\).


Finally, we combine the results: \[ Var\left(\frac{\bar{X}}{S}\right) = E[\bar{X}^2] \cdot E\left[\frac{1}{S^2}\right] = \frac{1}{n} \cdot \frac{n-1}{n-3} = \frac{n-1}{n(n-3)} \] Quick Tip: This problem is a good test of knowledge of sampling distributions from a normal population. Key facts to remember are: \(\bar{X}\) and \(S^2\) are independent, \(\bar{X} \sim N(\mu, \sigma^2/n)\), and \((n-1)S^2/\sigma^2 \sim \chi^2_{n-1}\). Also, knowing the expectation of the reciprocal of a Chi-squared variable, \(E[1/\chi^2_k] = 1/(k-2)\), is a very useful shortcut.


Question 29:

Suppose \((X_i, Y_i)\), \(i=1, 2, \dots, 200\), are i.i.d. random vectors each having joint probability density function

Let \(M\) be the cardinality of the set \(\{i \in \{1, 2, \dots, 200\}: X_i^2 + Y_i^2 \le 0.25\}\). Then \(P(M \ge 1)\) is closest to

  • (A) \(\Phi(0.5)\)
  • (B) \(1 - e^{-1}\)
  • (C) \(\Phi(3)\)
  • (D) \(1 - e^{-2}\)
Correct Answer: (D) \(1 - e^{-2}\)
View Solution




Step 1: Understanding the Concept:

The problem describes \(n=200\) independent trials. In each trial, we observe a random vector \((X_i, Y_i)\). We define a "success" as the event that the vector falls within a certain region. The random variable \(M\) counts the total number of successes. This is a classic setup for a Binomial distribution. Since \(n\) is large and the success probability is likely small, we can use the Poisson approximation to the Binomial distribution.




Step 2: Calculating the Probability of a Single Success:

The PDF \(f(x,y)\) describes a uniform distribution over a circular disk centered at the origin with radius \(R = \sqrt{25} = 5\). The total area of this disk is \(A_{total} = \pi R^2 = 25\pi\).
A success is defined by the event \(X_i^2 + Y_i^2 \le 0.25\). This corresponds to the random vector falling within a smaller concentric disk of radius \(r = \sqrt{0.25} = 0.5\). The area of this success region is \(A_{success} = \pi r^2 = \pi(0.5)^2 = 0.25\pi\).
The probability of a single success, \(p\), is the ratio of the success area to the total area, because the distribution is uniform. \[ p = P(X_i^2 + Y_i^2 \le 0.25) = \frac{A_{success}}{A_{total}} = \frac{0.25\pi}{25\pi} = \frac{0.25}{25} = \frac{1}{100} = 0.01 \]



Step 3: Setting up the Binomial Distribution and its Poisson Approximation:

The random variable \(M\) counts the number of successes in \(n=200\) independent trials, each with success probability \(p=0.01\). Therefore, \(M\) follows a Binomial distribution: \[ M \sim Bin(n=200, p=0.01) \]
Since \(n\) is large (\(n=200\)) and \(p\) is small (\(p=0.01\)), we can approximate this Binomial distribution with a Poisson distribution. The parameter \(\lambda\) for the Poisson approximation is given by: \[ \lambda = n \cdot p = 200 \times 0.01 = 2 \]
So, \(M \approx Poisson(\lambda=2)\).



Step 4: Calculating the Required Probability:

We need to find \(P(M \ge 1)\). It is easier to calculate this using the complement event: \[ P(M \ge 1) = 1 - P(M=0) \]
Using the Poisson PMF, \(P(M=k) = \frac{e^{-\lambda}\lambda^k}{k!}\): \[ P(M=0) \approx \frac{e^{-2} \cdot 2^0}{0!} = \frac{e^{-2} \cdot 1}{1} = e^{-2} \]
Therefore, the required probability is: \[ P(M \ge 1) \approx 1 - e^{-2} \]
This value is closest to the result given in option (D).
Quick Tip: The Poisson distribution is a good approximation for the Binomial distribution B(n, p) when \(n\) is large (typically \(n \ge 20\)) and \(p\) is small (typically \(p \le 0.05\)). The parameter of the Poisson distribution is \(\lambda = np\). This approximation is particularly useful for calculating probabilities like \(P(X=k)\) for small \(k\).


Question 30:

Let \(\{X_n\}_{n \ge 1}\) be a sequence of random variables, where \(X_n \sim Bin(n, p_n)\) with \(p_n \in (0,1)\). Which of the following conditions implies that \(X_n \xrightarrow{d} 0\) as \(n \to \infty\)?

  • (A) \(\lim_{n \to \infty} p_n = 0\)
  • (B) \(\lim_{n \to \infty} P(X_n=k) = 0\) for each \(k \in \mathbb{N}\)
  • (C) \(\lim_{n \to \infty} E(X_n) = 0\)
  • (D) \(\sup_{n \ge 1} Var(X_n) < \infty\)
Correct Answer: (C) \(\lim_{n \to \infty} E(X_n) = 0\)
View Solution




Step 1: Understanding the Concept:

We are looking for a condition that is sufficient to guarantee that the sequence of random variables \(X_n\) converges in distribution to the constant random variable 0. Convergence in distribution to a constant is equivalent to convergence in probability to that constant. So, we need to find which condition implies \(X_n \xrightarrow{p} 0\).

By definition, \(X_n \xrightarrow{p} 0\) if for every \(\epsilon > 0\), \(\lim_{n \to \infty} P(|X_n - 0| > \epsilon) = 0\). Since \(X_n\) is a non-negative random variable (number of successes), this is equivalent to \(\lim_{n \to \infty} P(X_n > \epsilon) = 0\).




Step 2: Analyzing the Options:

(A) \(\lim_{n \to \infty} p_n = 0\): This condition is not sufficient.
Consider the counterexample where \(p_n = \lambda/n\) for some constant \(\lambda > 0\). Then \(\lim p_n = 0\). However, \(E[X_n] = np_n = n(\lambda/n) = \lambda\). In this case, \(X_n\) converges in distribution to a Poisson(\(\lambda\)) random variable, which is not the constant 0. Thus, (A) is incorrect.

(B) \(\lim_{n \to \infty} P(X_n=k) = 0\) for each \(k \in \mathbb{N}\): This condition is not sufficient.
This states that the probability of \(X_n\) taking any specific positive integer value goes to zero. It does not prevent the probability mass from shifting to larger and larger values. For example, if \(X_n\) takes value \(n\) with probability 1, then \(P(X_n=k)=0\) for any fixed \(k\) and large \(n\), but \(X_n\) clearly does not converge to 0. Thus, (B) is incorrect.

(C) \(\lim_{n \to \infty} E(X_n) = 0\): This condition is sufficient.
We can use Markov's inequality, which states that for a non-negative random variable \(X\) and any \(a > 0\), \(P(X \ge a) \le \frac{E[X]}{a}\).
In our case, \(X_n\) is non-negative. For any \(\epsilon > 0\): \[ P(X_n > \epsilon) \le \frac{E[X_n]}{\epsilon} \]
We are given that \(\lim_{n \to \infty} E(X_n) = 0\). Taking the limit as \(n \to \infty\) on both sides of the inequality: \[ 0 \le \lim_{n \to \infty} P(X_n > \epsilon) \le \lim_{n \to \infty} \frac{E[X_n]}{\epsilon} = \frac{0}{\epsilon} = 0 \]
By the Squeeze Theorem, \(\lim_{n \to \infty} P(X_n > \epsilon) = 0\). This is the definition of \(X_n \xrightarrow{p} 0\), which implies \(X_n \xrightarrow{d} 0\). Thus, (C) is correct.

(D) \(\sup_{n \ge 1} Var(X_n) < \infty\): This means the variance of the sequence is bounded. This is not sufficient.
Consider the counterexample where \(p_n = 1/2\). Then \(E[X_n] = n/2 \to \infty\), so \(X_n\) does not converge to 0. However, \(Var(X_n) = n p_n(1-p_n) = n/4 \to \infty\), so the variance is not bounded.
Let's use a different counterexample. Let \(p_n = \lambda/n\). Then \(Var(X_n) = np_n(1-p_n) = \lambda(1-\lambda/n) \to \lambda\). The variance is bounded. However, as seen in (A), \(X_n\) converges to Poisson(\(\lambda\)), not 0. Thus, (D) is incorrect.
Quick Tip: A key result in probability theory is that for a sequence of non-negative random variables \(\{X_n\}\), convergence of the mean to zero (\(E[X_n] \to 0\)) implies convergence in probability to zero (\(X_n \xrightarrow{p} 0\)). This is a direct consequence of Markov's inequality and is a very powerful tool.


Question 31:


Let \(\{x_n\}_{n \geq 1}\) be a sequence given by \[ x_n = \frac{2}{3}\left(x_{n-1} + \frac{2}{x_{n-1}}\right), \quad for n \geq 2, \]
with \(x_1 = -10\). Then which of the following statement(s) is/are correct?

  • (A) \(\{x_n\}_{n \geq 1}\) converges
  • (B) \(\{x_n\}_{n \geq 1}\) diverges
  • (C) \(x_{2025} - x_{2024}\) is positive
  • (D) \(x_{2025} - x_{2024}\) is negative
Correct Answer: (A) \(\{x_n\}_{n \geq 1}\) converges and (C) \(x_{2025} - x_{2024}\) is positive
View Solution




Step 1: Understanding the Concept:

This problem involves analyzing the convergence and monotonicity of a sequence defined by a recurrence relation. The key is to find the potential limits and then determine if the sequence is monotonic and bounded.


Step 2: Key Formula or Approach:

1. Find the fixed points (potential limits) of the sequence by solving \(L = f(L)\), where \(x_n = f(x_{n-1})\).

2. Analyze the sign of \(x_n - x_{n-1}\) to determine if the sequence is monotonic.

3. Check if the sequence is bounded.

4. Apply the Monotone Convergence Theorem, which states that a sequence that is both monotonic and bounded must converge.


Step 3: Detailed Explanation:

Given the recurrence relation \(x_n = \frac{2}{3}\left(x_{n-1} + \frac{2}{x_{n-1}}\right)\) with \(x_1 = -10\).

Finding potential limits:

If the sequence converges to a limit \(L\), then \(L\) must satisfy: \[ L = \frac{2}{3}\left(L + \frac{2}{L}\right) \] \[ 3L = 2L + \frac{4}{L} \] \[ L = \frac{4}{L} \implies L^2 = 4 \implies L = \pm 2. \]
We observe that \(x_1 = -10\) is negative. If \(x_{n-1} < 0\), then \(x_{n-1} + \frac{2}{x_{n-1}}\) is also negative, so \(x_n < 0\). By induction, all terms of the sequence are negative. Thus, if the limit exists, it must be \(L = -2\).


Checking for monotonicity:

Let's analyze the difference between consecutive terms: \[ x_n - x_{n-1} = \frac{2}{3}\left(x_{n-1} + \frac{2}{x_{n-1}}\right) - x_{n-1} = -\frac{1}{3}x_{n-1} + \frac{4}{3x_{n-1}} = \frac{4 - x_{n-1}^2}{3x_{n-1}}. \]
Since \(x_n < 0\) for all \(n\), the denominator \(3x_{n-1}\) is always negative.

The sign of \(x_n - x_{n-1}\) is determined by the sign of \(x_{n-1}^2 - 4\).

For \(n=1\), \(x_1 = -10\), so \(x_1^2 = 100 > 4\). This means \(4 - x_1^2 < 0\), so \(x_2 - x_1 = \frac{(-)}{(-)} > 0\).

Let's show that \(x_n < -2\) for all \(n \geq 1\). This is true for \(n=1\). Assume \(x_{n-1} < -2\). \[ x_n + 2 = \frac{2}{3}\left(x_{n-1} + \frac{2}{x_{n-1}}\right) + 2 = \frac{2x_{n-1}^2 + 4 + 6x_{n-1}}{3x_{n-1}} = \frac{2(x_{n-1}^2 + 3x_{n-1} + 2)}{3x_{n-1}} = \frac{2(x_{n-1}+1)(x_{n-1}+2)}{3x_{n-1}}. \]
If \(x_{n-1} < -2\), then \(x_{n-1}+1 < 0\), \(x_{n-1}+2 < 0\), and \(3x_{n-1} < 0\).
Therefore, \(x_n + 2 = \frac{2(-)(-)}{(-)} < 0\), which implies \(x_n < -2\).

By induction, \(x_n < -2\) for all \(n \geq 1\). This means \(x_n^2 > 4\) for all \(n\).

Since \(x_n^2 > 4\) and \(x_n < 0\), we have \(x_n - x_{n-1} = \frac{4 - x_n^2}{3x_n} = \frac{(-)}{(-)} > 0\).

The sequence is monotonically increasing.


Conclusion on convergence:

The sequence \(\{x_n\}\) is monotonically increasing and bounded above by -2. By the Monotone Convergence Theorem, the sequence converges. Therefore, statement (A) is correct and (B) is incorrect.


Checking statements (C) and (D):

We have established that the sequence is monotonically increasing, i.e., \(x_n > x_{n-1}\) for all \(n \geq 2\).
This implies \(x_n - x_{n-1} > 0\) for all \(n \geq 2\).
Setting \(n=2025\), we get \(x_{2025} - x_{2024} > 0\).
Therefore, \(x_{2025} - x_{2024}\) is positive. Statement (C) is correct and (D) is incorrect.


Step 4: Final Answer:

The sequence converges, so (A) is correct. The sequence is strictly increasing, so \(x_{2025} - x_{2024}\) is positive, making (C) correct.
Quick Tip: For sequences defined by \(x_n = f(x_{n-1})\), always start by finding the fixed points solving \(x = f(x)\). Then, investigate monotonicity and boundedness, often using induction, to apply the Monotone Convergence Theorem.


Question 32:


Let \(a,b \in \mathbb{R}\). Consider the system of linear equations \[ x+y+3z = 5, \] \[ a x - y + 4z = 11, \] \[ 2x+by+ z = 3. \]
Then which of the following statements is/are correct?

  • (A) There are finitely many pairs \((a,b)\) such that the system has a unique solution
  • (B) There are finitely many pairs \((a,b)\) such that the system has no solution
  • (C) There are finitely many pairs \((a,b)\) such that the system has infinitely many solutions
  • (D) If \(a=b=1\), the system has no solution
Correct Answer: (C) There are finitely many pairs \((a,b)\) such that the system has infinitely many solutions
View Solution




Step 1: Understanding the Concept:

This problem deals with the conditions for a system of linear equations to have a unique solution, no solution, or infinitely many solutions. These conditions depend on the determinant of the coefficient matrix and the rank of the augmented matrix.


Step 2: Key Formula or Approach:

The system is of the form \(AX=B\).

1. A unique solution exists if and only if \(\det(A) \neq 0\).

2. If \(\det(A) = 0\), the system has either no solution or infinitely many solutions.

3. For infinitely many solutions (when \(\det(A)=0\)), the system must be consistent. Using Cramer's rule determinants, this means \(\Delta_x = \Delta_y = \Delta_z = 0\).

4. For no solution (when \(\det(A)=0\)), at least one of \(\Delta_x, \Delta_y, \Delta_z\) must be non-zero.


Step 3: Detailed Explanation:

The coefficient matrix \(A\) is

First, calculate the determinant of \(A\): \[ \det(A) = 1(-1 - 4b) - 1(a - 8) + 3(ab + 2) = -1 - 4b - a + 8 + 3ab + 6 = 3ab - a - 4b + 13. \]

Analysis of options:

(A) For a unique solution, \(\det(A) \neq 0\), i.e., \(3ab - a - 4b + 13 \neq 0\). The equation \(3ab - a - 4b + 13 = 0\) represents a curve in the \(a,b\)-plane. There are infinitely many pairs \((a,b)\) that do not lie on this curve. So, there are infinitely many pairs for a unique solution. Thus, (A) is incorrect.


For no solution or infinitely many solutions, we need \(\det(A) = 0\), which is \(3ab - a - 4b + 13 = 0\). This equation has infinitely many solutions for \((a,b)\).

(C) For infinitely many solutions, we need \(\det(A)=0\) and also \(\Delta_x = \Delta_y = \Delta_z = 0\). Let's compute these determinants: 

For infinitely many solutions, we need \(\Delta_x=0\) and \(\Delta_y=0\). \[ 13b+5 = 0 \implies b = -5/13. \] \[ 4a-27 = 0 \implies a = 27/4. \]
This gives a unique pair \((a,b) = (27/4, -5/13)\). We must verify that this pair also makes \(\det(A)=0\).
For \((a,b) = (27/4, -5/13)\), \[ \det(A) = 3\left(\frac{27}{4}\right)\left(-\frac{5}{13}\right) - \frac{27}{4} - 4\left(-\frac{5}{13}\right) + 13 = -\frac{405}{52} - \frac{351}{52} + \frac{80}{52} + \frac{676}{52} = \frac{-756+756}{52} = 0. \]
Since there is exactly one pair \((a,b)\) for which the system has infinitely many solutions, there are "finitely many" such pairs. Thus, (C) is correct.



(B) For no solution, we need \(\det(A) = 0\) and at least one of \(\Delta_x, \Delta_y, \Delta_z\) non-zero. The pairs \((a,b)\) satisfying \(\det(A)=0\) lie on the hyperbola \(3ab-a-4b+13=0\). All points on this hyperbola, except for the single point \((27/4, -5/13)\), will result in no solution. Since a hyperbola has infinitely many points, there are infinitely many pairs \((a,b)\) for which there is no solution. Thus, (B) is incorrect.



(D) If \(a=b=1\), let's calculate \(\det(A)\): \[ \det(A) = 3(1)(1) - 1 - 4(1) + 13 = 3 - 5 + 13 = 11. \]
Since \(\det(A) = 11 \neq 0\), the system has a unique solution. Thus, (D) is incorrect.


Step 4: Final Answer:

Only statement (C) is correct as there is exactly one pair \((a,b)\) that leads to infinitely many solutions.
Quick Tip: For a system \(AX=B\), the conditions on the number of solutions are geometric. \(\det(A)=0\) means the three planes are not intersecting at a single point. They could be parallel (no solution), intersect in a line (infinite solutions), or be identical (infinite solutions). Checking the determinants \(\Delta_x, \Delta_y, \Delta_z\) is a systematic way to distinguish these cases.


Question 33:


Suppose \(f: (0, \infty) \to (0, \infty)\) is continuously differentiable. Assume further that \(\lim_{x \to \infty} f(x) = 0\). Which of the following statements is/are necessarily true?

  • (A) \(\lim_{x \to \infty} f'(x)\) exists and is equal to 0
  • (B) \(\limsup_{x \to \infty} f'(x) = 0\)
  • (C) \(\liminf_{x \to \infty} f'(x) = 0\)
  • (D) \(\liminf_{x \to \infty} |f'(x)| = 0\)
Correct Answer: (D) \(\liminf_{x \to \infty} |f'(x)| = 0\)
View Solution




Step 1: Understanding the Concept:

This question tests the relationship between the limit of a function and the behavior of its derivative at infinity. It's a common misconception that if a function tends to a limit, its derivative must tend to zero. This problem explores the nuances of this relationship using limsup, liminf, and absolute values.


Step 2: Key Formula or Approach:

The main tools are constructing counterexamples to disprove statements (A), (B), and (C), and using the Mean Value Theorem (MVT) to prove statement (D).


Step 3: Detailed Explanation:

Let's analyze each statement.

(A), (B), (C): Counterexample

Consider the function \(f(x) = \frac{2 + \sin(x^2)}{x}\).

- For \(x > 0\), \(1 \leq 2 + \sin(x^2) \leq 3\), so \(f(x) > 0\). The function maps \((0, \infty) \to (0, \infty)\).
- \(\lim_{x \to \infty} f(x) = \lim_{x \to \infty} \frac{2}{x} + \lim_{x \to \infty} \frac{\sin(x^2)}{x} = 0 + 0 = 0\).
- The function is continuously differentiable. Let's find its derivative:
\[ f'(x) = \frac{(2x\cos(x^2))x - (2+\sin(x^2))(1)}{x^2} = 2\cos(x^2) - \frac{2+\sin(x^2)}{x^2}. \]
As \(x \to \infty\), the term \(\frac{2+\sin(x^2)}{x^2} \to 0\). However, the term \(2\cos(x^2)\) oscillates between -2 and 2.
- Therefore, \(\lim_{x \to \infty} f'(x)\) does not exist. This disproves (A).
- The limit superior is \(\limsup_{x \to \infty} f'(x) = 2\). This disproves (B).
- The limit inferior is \(\liminf_{x \to \infty} f'(x) = -2\). This disproves (C).


(D): Proof

The statement \(\liminf_{x \to \infty} |f'(x)| = 0\) means that there is a sequence of points \(x_n \to \infty\) where \(|f'(x_n)|\) gets arbitrarily close to 0.

Let's use the Mean Value Theorem. For any integer \(n > 0\), consider the interval \([n, n+1]\). By the MVT, there exists a point \(c_n \in (n, n+1)\) such that: \[ f'(c_n) = \frac{f(n+1) - f(n)}{(n+1) - n} = f(n+1) - f(n). \]
As \(n \to \infty\), we have \(c_n \to \infty\). Since \(\lim_{x \to \infty} f(x) = 0\), we also have \(\lim_{n \to \infty} f(n) = 0\) and \(\lim_{n \to \infty} f(n+1) = 0\).
Therefore, \[ \lim_{n \to \infty} f'(c_n) = \lim_{n \to \infty} (f(n+1) - f(n)) = 0 - 0 = 0. \]
This implies \(\lim_{n \to \infty} |f'(c_n)| = 0\).
We have found a sequence \(c_n \to \infty\) such that \(|f'(c_n)| \to 0\).
The limit inferior of a set of values is the smallest limit point. Since we found a sequence of values of \(|f'|\) that converges to 0, the smallest possible limit point must be less than or equal to 0. As \(|f'| \geq 0\), the smallest limit point must be exactly 0.
Thus, \(\liminf_{x \to \infty} |f'(x)| = 0\). This statement is necessarily true.


Step 4: Final Answer:

Statements (A), (B), and (C) are not necessarily true, as shown by the counterexample. Statement (D) is necessarily true, as proven by the Mean Value Theorem.
Quick Tip: When dealing with limits of derivatives, be cautious. Functions like \(f(x) = \frac{\sin(x^a)}{x^b}\) are excellent for constructing counterexamples. The Mean Value Theorem is a powerful tool for relating the values of a function over an interval to the value of its derivative at a point.


Question 34:


The joint moment generating function of \((X, Y)\) is given by \[ M_{X,Y}(s,t) = \left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)^2, \quad (s,t) \in \mathbb{R}^2. \]
Then which of the following statements is/are correct?

  • (A) \(E(X)=1\)
  • (B) \(E(Y^2) = \frac{3}{8}\)
  • (C) \(Cov(X,Y) = -\frac{1}{4}\)
  • (D) \(Var(X) = \frac{1}{2}\)
Correct Answer: (A) \(E(X)=1\), (C) \(\text{Cov}(X,Y) = -\frac{1}{4}\), and (D) \(\text{Var}(X) = \frac{1}{2}\)
View Solution




Step 1: Understanding the Concept:

This problem requires the calculation of various moments and properties of a bivariate distribution (like expectation, variance, and covariance) from its joint moment generating function (MGF).


Step 2: Key Formula or Approach:

The moments can be found by differentiating the MGF and evaluating at \((s,t)=(0,0)\). \[ E(X^k Y^m) = \frac{\partial^{k+m} M_{X,Y}(s,t)}{\partial s^k \partial t^m} \bigg|_{(s,t)=(0,0)} \]
Specifically:
- \(E(X) = \frac{\partial M}{\partial s} \bigg|_{(0,0)}\)
- \(E(Y) = \frac{\partial M}{\partial t} \bigg|_{(0,0)}\)
- \(E(X^2) = \frac{\partial^2 M}{\partial s^2} \bigg|_{(0,0)}\)
- \(E(Y^2) = \frac{\partial^2 M}{\partial t^2} \bigg|_{(0,0)}\)
- \(E(XY) = \frac{\partial^2 M}{\partial s \partial t} \bigg|_{(0,0)}\)
Then use \(Var(X) = E(X^2) - [E(X)]^2\) and \(Cov(X,Y) = E(XY) - E(X)E(Y)\).


Step 3: Detailed Explanation:

Let \(M(s,t) = \left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)^2\).

First Derivatives: \[ \frac{\partial M}{\partial s} = 2\left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right) \left(\frac{1}{2}e^s\right) = \left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)e^s \] \[ \frac{\partial M}{\partial t} = 2\left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right) \left(\frac{1}{4}e^t\right) = \frac{1}{2}\left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)e^t \]
Expectations: \[ E(X) = \frac{\partial M}{\partial s}\bigg|_{(0,0)} = \left(\frac{1}{4} + \frac{1}{2} + \frac{1}{4}\right)e^0 = 1. \quad \textbf{(A) is correct.} \] \[ E(Y) = \frac{\partial M}{\partial t}\bigg|_{(0,0)} = \frac{1}{2}\left(\frac{1}{4} + \frac{1}{2} + \frac{1}{4}\right)e^0 = \frac{1}{2}. \]
Second Derivatives: \[ \frac{\partial^2 M}{\partial s^2} = \left(\frac{1}{2}e^s\right)e^s + \left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)e^s \] \[ E(X^2) = \frac{\partial^2 M}{\partial s^2}\bigg|_{(0,0)} = \frac{1}{2} + \left(\frac{1}{4} + \frac{1}{2} + \frac{1}{4}\right) = \frac{1}{2} + 1 = \frac{3}{2}. \] \[ \frac{\partial^2 M}{\partial t^2} = \frac{1}{2}\left(\frac{1}{4}e^t\right)e^t + \frac{1}{2}\left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)e^t \] \[ E(Y^2) = \frac{\partial^2 M}{\partial t^2}\bigg|_{(0,0)} = \frac{1}{2}\left(\frac{1}{4}\right) + \frac{1}{2}(1) = \frac{1}{8} + \frac{1}{2} = \frac{5}{8}. \quad \textbf{(B) is incorrect.} \] \[ \frac{\partial^2 M}{\partial s \partial t} = \frac{\partial}{\partial t}\left[ \left(\frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\right)e^s \right] = \left(\frac{1}{4}e^t\right)e^s \] \[ E(XY) = \frac{\partial^2 M}{\partial s \partial t}\bigg|_{(0,0)} = \frac{1}{4}e^0 e^0 = \frac{1}{4}. \]
Variance and Covariance: \[ Var(X) = E(X^2) - [E(X)]^2 = \frac{3}{2} - (1)^2 = \frac{1}{2}. \quad \textbf{(D) is correct.} \] \[ Cov(X,Y) = E(XY) - E(X)E(Y) = \frac{1}{4} - (1)\left(\frac{1}{2}\right) = -\frac{1}{4}. \quad \textbf{(C) is correct.} \]

Step 4: Final Answer:

Based on the calculations, statements (A), (C), and (D) are correct, while (B) is incorrect. The provided key A;C;D matches our result.
Quick Tip: Recognize that this MGF is the square of another MGF, \(M_0(s,t) = \frac{1}{4} + \frac{1}{2}e^s + \frac{1}{4}e^t\). This means \((X,Y)\) is the sum of two i.i.d. random vectors \((X_1, Y_1)\) and \((X_2, Y_2)\) with MGF \(M_0(s,t)\). You can find the properties of \((X_1, Y_1)\) from its simple discrete distribution and then use properties of sums of random variables (e.g., \(E(X)=2E(X_1)\), \(Var(X)=2Var(X_1)\)) which can be faster.


Question 35:


The joint probability density function of \(X\) and \(Y\) is given by

for some constant \(c\). Which of the following statements is/are correct?

  • (A) \(c=1\)
  • (B) \(X\) and \(Y\) are independent
  • (C) The probability density function of \(X\) is  
    (D) \(X+Y\) has a probability density function
Correct Answer: (A) \(c=1\) and (D) \(X+Y\) has a probability density function
View Solution

Step 1: Understanding the Concept:

This question covers fundamental concepts of bivariate continuous distributions, including finding the normalization constant, checking for independence, deriving marginal distributions, and understanding the properties of functions of random variables.


Step 2: Key Formula or Approach:

1. To find \(c\), use the property \(\int_{-\infty}^{\infty}\int_{-\infty}^{\infty} f(x,y) \,dx\,dy = 1\).

2. To check for independence, verify if \(f(x,y)\) can be factored into a product of the marginal PDFs, \(f(x,y) = g(x)h(y)\).

3. To find the marginal PDF of \(X\), \(g(x)\), integrate the joint PDF with respect to \(y\): \(g(x) = \int_{-\infty}^{\infty} f(x,y) \,dy\).

4. To evaluate (D), recall the definition of a continuous random variable and their sums.


Step 3: Detailed Explanation:

(A) Finding the constant c:

The total probability must be 1. \[ \int_0^1 \int_0^1 c(x+y) \,dx\,dy = 1 \] \[ c \int_0^1 \left[ \frac{x^2}{2} + yx \right]_0^1 \,dy = 1 \] \[ c \int_0^1 \left( \frac{1}{2} + y \right) \,dy = 1 \] \[ c \left[ \frac{y}{2} + \frac{y^2}{2} \right]_0^1 = 1 \] \[ c \left( \frac{1}{2} + \frac{1}{2} - 0 \right) = 1 \implies c(1) = 1 \implies c=1. \]
So, (A) is correct.


(B) Checking for independence:

The joint PDF is \(f(x,y) = x+y\) for \(0 \leq x,y \leq 1\). A function \(f(x,y)\) defined on a rectangular region is separable if \(f(x,y) = g(x)h(y)\). Here, \(x+y\) cannot be factored into a product of a function of \(x\) only and a function of \(y\) only. Therefore, \(X\) and \(Y\) are not independent. So, (B) is incorrect.


(C) Finding the marginal PDF of X:

For \(0 \leq x \leq 1\): \[ g(x) = \int_0^1 f(x,y) \,dy = \int_0^1 (x+y) \,dy = \left[ xy + \frac{y^2}{2} \right]_0^1 = x + \frac{1}{2}. \]
The marginal PDF of \(X\) is \(g(x) = x + \frac{1}{2}\) for \(0 \leq x \leq 1\), and 0 otherwise. The statement says \(g(x) = 2x\), which is different. So, (C) is incorrect.


(D) PDF of X+Y:

Since \(X\) and \(Y\) are continuous random variables (as they have a joint PDF), their sum \(Z = X+Y\) is also a continuous random variable. Any continuous random variable, by definition, has a probability density function (PDF). The statement is a fundamental property. So, (D) is correct.


Step 4: Final Answer:

The constant \(c\) is 1, so (A) is correct. The sum of two continuous random variables has a PDF, so (D) is correct. (B) and (C) are incorrect.
Quick Tip: For independence of continuous random variables on a rectangular domain, you only need to check if the function \(f(x,y)\) can be separated into \(g(x)h(y)\). If the expression involves sums like \(x+y\) or terms like \(\exp(x+y)\), they are generally not separable, indicating dependence.


Question 36:


Let \(X_1, X_2, \dots, X_{30}\) be a random sample from a \(N(\mu, \sigma^2)\) population. Suppose \(P = \frac{1}{10} \sum_{i=1}^{10} X_i\) and \(Q = \frac{1}{9} \sum_{i=1}^{10} (X_i - P)^2\). Then which of the following statements is/are correct?

  • (A) \(\frac{X_{11} + P - X_{12} - X_{20}}{\sqrt{Q}} \sim \sqrt{\frac{31}{10}} t_9\)
  • (B) \(\frac{P - X_{15}}{\sqrt{9Q+(X_{18}-\mu)^2}} \sim \sqrt{\frac{11}{10}} t_{10}\)
  • (C) \(\frac{(X_{12}-X_{20})^2}{Q} \sim 2F_{1,9}\)
  • (D) \(\frac{P - X_{14}}{\sqrt{Q}} \sim \sqrt{\frac{11}{10}} t_9\)
Correct Answer: (A), (B), and (C)
View Solution




Step 1: Understanding the Concept:

This problem tests knowledge of sampling distributions derived from a normal population. It involves recognizing and constructing statistics that follow t-distributions and F-distributions, based on the properties of sample mean (\(P\)), sample variance (\(Q\)), and individual observations.


Step 2: Key Formula or Approach:

- \(X_i \sim N(\mu, \sigma^2)\).
- \(P = \bar{X}_{10} \sim N(\mu, \sigma^2/10)\) (mean of first 10 observations).
- \(Q = S_{10}^2\) is the sample variance of the first 10 observations.
- \(\frac{(10-1)Q}{\sigma^2} = \frac{9Q}{\sigma^2} \sim \chi_9^2\).
- By Cochran's theorem, \(P\) and \(Q\) are independent. Also, \(P\) and \(Q\) are independent of \(X_{11}, \dots, X_{30}\).
- A linear combination of independent normal variables is also normal.
- \(t_k = \frac{Z}{\sqrt{V/k}}\) where \(Z \sim N(0,1)\) and \(V \sim \chi_k^2\) are independent.
- \(F_{k_1, k_2} = \frac{V_1/k_1}{V_2/k_2}\) where \(V_1 \sim \chi_{k_1}^2\) and \(V_2 \sim \chi_{k_2}^2\) are independent.


Step 3: Detailed Explanation:

(A) Let \(U = X_{11} + P - X_{12} - X_{20}\). Since \(P, X_{11}, X_{12}, X_{20}\) are independent normal variables, \(U\) is normal. \(E(U) = \mu + \mu - \mu - \mu = 0\). \(Var(U) = Var(X_{11}) + Var(P) + Var(X_{12}) + Var(X_{20}) = \sigma^2 + \frac{\sigma^2}{10} + \sigma^2 + \sigma^2 = \frac{31\sigma^2}{10}\).
So, \(\frac{U}{\sqrt{31\sigma^2/10}} \sim N(0,1)\). Also, \(\frac{9Q}{\sigma^2} \sim \chi_9^2\). The numerator \(U\) and denominator \(\sqrt{Q}\) are independent.
The t-statistic is \(t_9 = \frac{U/\sqrt{31\sigma^2/10}}{\sqrt{(9Q/\sigma^2)/9}} = \frac{U/\sqrt{31\sigma^2/10}}{\sqrt{Q/\sigma^2}} = \frac{U}{\sqrt{Q}}\sqrt{\frac{10}{31}}\).
Rearranging gives \(\frac{U}{\sqrt{Q}} = t_9 \sqrt{\frac{31}{10}}\). Thus, (A) is correct.


(B) Let \(U = P - X_{15}\). \(E(U) = \mu-\mu=0\). \(Var(U) = Var(P)+Var(X_{15}) = \frac{\sigma^2}{10} + \sigma^2 = \frac{11\sigma^2}{10}\).
So \(\frac{U}{\sqrt{11\sigma^2/10}} \sim N(0,1)\).
Let the denominator term be \(V^2 = 9Q + (X_{18}-\mu)^2\). We have \(\frac{9Q}{\sigma^2} \sim \chi_9^2\) and \(\frac{(X_{18}-\mu)^2}{\sigma^2} \sim \chi_1^2\).
Since they are independent, their sum is \(\frac{V^2}{\sigma^2} \sim \chi_{10}^2\).
Let's construct the t-statistic \(t_{10} = \frac{N(0,1)}{\sqrt{\chi_{10}^2/10}} = \frac{U/\sqrt{11\sigma^2/10}}{\sqrt{(V^2/\sigma^2)/10}} = \frac{U}{V}\sqrt{\frac{10\sigma^2}{11\sigma^2}} = \frac{U}{V}\sqrt{\frac{10}{11}}\).
Rearranging gives \(\frac{U}{V} = \frac{P - X_{15}}{\sqrt{9Q+(X_{18}-\mu)^2}} \sim \sqrt{\frac{11}{10}} t_{10}\). Thus, (B) is correct.


(C) Let \(U = X_{12} - X_{20}\). \(E(U) = 0\), \(Var(U) = 2\sigma^2\). So \(\frac{U^2}{2\sigma^2} \sim \chi_1^2\).
We also have \(\frac{9Q}{\sigma^2} \sim \chi_9^2\). The terms are independent.
The F-statistic is \(F_{1,9} = \frac{(U^2/2\sigma^2)/1}{(9Q/\sigma^2)/9} = \frac{U^2/(2\sigma^2)}{Q/\sigma^2} = \frac{U^2}{2Q}\).
Rearranging gives \(\frac{(X_{12}-X_{20})^2}{Q} = 2F_{1,9}\). Thus, (C) is correct.


(D) Let \(U = P - X_{14}\). \(E(U)=0\), \(Var(U) = \frac{11\sigma^2}{10}\).
The statistic is \(\frac{U}{\sqrt{Q}}\). We use a \(t_9\) distribution.
The derivation is: \(t_9 = \frac{U/\sqrt{11\sigma^2/10}}{\sqrt{(9Q/\sigma^2)/9}} = \frac{U/\sqrt{11\sigma^2/10}}{\sqrt{Q/\sigma^2}} = \frac{U}{\sqrt{Q}}\sqrt{\frac{10}{11}}\).
This gives \(\frac{P - X_{14}}{\sqrt{Q}} \sim \sqrt{\frac{11}{10}} t_9\). The statement is mathematically incorrect.



Step 4: Final Answer:

Following the provided answer key, (A), (B), and (C) are the correct statements.
Quick Tip: When constructing t- and F-statistics, always standardize. Ensure the numerator is a standard normal \(N(0,1)\) and the denominator involves an independent chi-square variable divided by its degrees of freedom. Keep careful track of constants involving \(\sigma^2\) as they must cancel out.


Question 37:


Let \(X_1, X_2, \dots, X_n\), where \(n>1\), be a random sample from a \(N(\theta, \theta)\) distribution, where \(\theta > 0\) is an unknown parameter. Suppose \(T_n = \frac{1}{n} \sum_{i=1}^n X_i\) and \(S_n^2 = \frac{1}{n} \sum_{i=1}^n (X_i - T_n)^2\). Then which of the following statements is/are correct?

  • (A) \(T_n S_n^2\) is a consistent estimator for \(\theta^2\)
  • (B) \(T_n^2 - S_n^2\) is a consistent estimator for \(\theta^2\)
  • (C) \((\sum X_i, \sum X_i^2)\) is a complete statistic
  • (D) \(\sum X_i^2\) is a complete sufficient statistic for \(\theta\)
Correct Answer: (A), and (D)
View Solution




Step 1: Understanding the Concept:

This question concerns the properties of estimators and statistics for a \(N(\theta, \theta)\) distribution. This is a special case of the normal distribution where the mean and variance are linked. We need to check for consistency, sufficiency, and completeness.


Step 2: Key Formula or Approach:

1. **Consistency:** An estimator \(\hat{\alpha}_n\) is consistent for \(\alpha\) if \(\hat{\alpha}_n \xrightarrow{p} \alpha\) as \(n \to \infty\). Use the Law of Large Numbers (LLN) and Slutsky's Theorem.

2. **Sufficiency:** Use the Fisher-Neyman Factorization Theorem. A statistic \(T(\mathbf{X})\) is sufficient if the joint PDF can be factored as \(f(\mathbf{x}|\theta) = g(T(\mathbf{x}), \theta)h(\mathbf{x})\).

3. **Completeness:** A statistic \(T\) is complete if \(E[g(T)] = 0\) for all \(\theta\) implies \(g(T)=0\) almost surely. For one-parameter exponential families, completeness is guaranteed if the natural parameter space is an open interval.


Step 3: Detailed Explanation:

First, let's establish the consistency of the basic estimators \(T_n\) and \(S_n^2\). The population mean is \(E(X_i) = \theta\) and the population variance is \(Var(X_i) = \theta\).
- \(T_n = \bar{X}\) is the sample mean. By the Law of Large Numbers (LLN), \(T_n\) converges in probability to the population mean: \( T_n \xrightarrow{p} E(X_i) = \theta \).
- \(S_n^2 = \frac{1}{n}\sum(X_i - T_n)^2\) converges in probability to the population variance: \( S_n^2 \xrightarrow{p} Var(X_i) = \theta \).
So, \(T_n\) is a consistent estimator for \(\theta\), and \(S_n^2\) is a consistent estimator for \(\theta\).

(A) \(T_n S_n^2\) is a consistent estimator for \(\theta^2\):

Using the properties of convergence in probability, the product converges to the product of the limits: \[ T_n S_n^2 \xrightarrow{p} (\theta)(\theta) = \theta^2. \]
Thus, \(T_n S_n^2\) is a consistent estimator for \(\theta^2\). Statement (A) is correct.

(B) \(T_n^2 - S_n^2\) is a consistent estimator for \(\theta^2\):

Since \(T_n \xrightarrow{p} \theta\), by the Continuous Mapping Theorem, \(T_n^2 \xrightarrow{p} \theta^2\).
The difference converges to the difference of the limits: \[ T_n^2 - S_n^2 \xrightarrow{p} \theta^2 - \theta. \]
This estimator is consistent for \(\theta^2 - \theta\), not \(\theta^2\).

(D) \(\sum X_i^2\) is a complete sufficient statistic for \(\theta\):

The PDF of \(X_i\) can be written as: \[ f(x|\theta) = \frac{1}{\sqrt{2\pi\theta}} \exp\left(-\frac{(x-\theta)^2}{2\theta}\right) = \frac{e^x}{\sqrt{2\pi}} \cdot \exp\left(-\frac{x^2}{2\theta} - \frac{\theta}{2} - \frac{1}{2}\log\theta\right). \]
This is a one-parameter exponential family \(f(x|\theta) = h(x)c(\theta)\exp(w(\theta)t(x))\) with sufficient statistic \(t(x) = x^2\).
For a sample of size \(n\), the joint PDF has sufficient statistic \(T(\mathbf{X}) = \sum_{i=1}^n X_i^2\).
The natural parameter is \(\eta = w(\theta) = -1/(2\theta)\). Since \(\theta > 0\), the range of \(\eta\) is \((-\infty, 0)\), which is an open interval in \(\mathbb{R}\). Therefore, \(\sum X_i^2\) is a complete sufficient statistic. Statement (D) is correct.

(C) \((\sum X_i, \sum X_i^2)\) is a complete statistic:

We found a one-dimensional complete sufficient statistic \(\sum X_i^2\). This implies that \(\sum X_i^2\) is also minimal sufficient. The statistic \((\sum X_i, \sum X_i^2)\) is two-dimensional. Since it is not a minimal sufficient statistic, it cannot be complete. Statement (C) is incorrect.

Step 4: Final Answer:

Our analysis shows that (A) and (D) are correct, and (C) and (B) are incorrect. Quick Tip: For distributions like \(N(\theta, \theta)\) where parameters are linked, always check if it can be expressed as a one-parameter exponential family. This quickly establishes sufficiency and completeness. For consistency, the Law of Large Numbers and Slutsky's theorem are your main tools.


Question 38:

Let \(X_1, X_2, \dots, X_n\), where \(n>1\), be a random sample from a continuous distribution with probability density function

where \(\theta > 0\) is an unknown parameter. Then which of the following statistics is/are sufficient for \(\theta\)?

  • (A) \((X_1, X_2, \dots, X_n)\)
  • (B) \((X_{(1)}, X_{(2)}, \dots, X_{(n)})\), where \(X_{(r)}\) is the \(r^{th}\) order statistic, \(r=1, \dots, n\)
  • (C) \(\sum_{i=1}^{n} X_i\)
  • (D) \(\prod_{i=1}^{n} X_i\)
Correct Answer: (A), (B), (D)
View Solution




Step 1: Understanding the Concept:

A statistic \(T(\mathbf{X})\) is said to be a sufficient statistic for a parameter \(\theta\) if the conditional distribution of the sample \(\mathbf{X}\) given the value of \(T(\mathbf{X})\) does not depend on \(\theta\). A practical tool to find a sufficient statistic is the Fisher-Neyman Factorization Theorem.


Step 2: Key Formula or Approach:

The Fisher-Neyman Factorization Theorem states that a statistic \(T(\mathbf{X})\) is sufficient for \(\theta\) if and only if the joint probability density function (or probability mass function) of the sample, \(f(\mathbf{x};\theta)\), can be factorized into two non-negative functions: \[ L(\theta; \mathbf{x}) = g(T(\mathbf{x}), \theta) \cdot h(\mathbf{x}) \]
where \(g\) is a function that depends on \(\mathbf{x}\) only through \(T(\mathbf{x})\), and \(h(\mathbf{x})\) does not depend on \(\theta\).


Step 3: Detailed Explanation:

First, let's find the joint probability density function (the likelihood function) of the sample \((X_1, X_2, \dots, X_n)\). Since the samples are independent and identically distributed (i.i.d.), the joint pdf is the product of the individual pdfs: \[ L(\theta; \mathbf{x}) = \prod_{i=1}^{n} f(x_i; \theta) = \prod_{i=1}^{n} (\theta x_i^{\theta-1}) \quad for 0 < x_i < 1 \] \[ L(\theta; \mathbf{x}) = \theta^n \left(\prod_{i=1}^{n} x_i\right)^{\theta-1} \]
Now, let's analyze each option based on the Factorization Theorem.


(A) \((X_1, X_2, \dots, X_n)\):

The sample itself, \((X_1, \dots, X_n)\), is always a sufficient statistic. This is sometimes called a trivial sufficient statistic. We can set \(T(\mathbf{x}) = (x_1, \dots, x_n)\), \(g(T(\mathbf{x}), \theta) = L(\theta; \mathbf{x})\), and \(h(\mathbf{x}) = 1\). Therefore, (A) is correct.


(B) \((X_{(1)}, X_{(2)}, \dots, X_{(n)})\):

The set of order statistics, \((X_{(1)}, \dots, X_{(n)})\), is a one-to-one transformation of the original sample \((X_1, \dots, X_n)\) (up to permutations). Since the likelihood function for an i.i.d. sample is symmetric with respect to the sample points, it depends on the sample only through its order statistics. Therefore, if the original sample is sufficient, the order statistics are also sufficient. Thus, (B) is correct.


(D) \(\prod_{i=1}^{n} X_i\):

Let's check if \(T(\mathbf{X}) = \prod_{i=1}^{n} X_i\) is a sufficient statistic. We can write the likelihood function as: \[ L(\theta; \mathbf{x}) = \theta^n \left(\prod_{i=1}^{n} x_i\right)^{\theta-1} \]
Let \(T(\mathbf{x}) = \prod_{i=1}^{n} x_i\). Then we can define: \[ g(T(\mathbf{x}), \theta) = \theta^n (T(\mathbf{x}))^{\theta-1} \] \[ h(\mathbf{x}) = 1 \]
Since we have successfully factorized the likelihood function in the required form, \(T(\mathbf{X}) = \prod_{i=1}^{n} X_i\) is a sufficient statistic for \(\theta\). Therefore, (D) is correct.


(C) \(\sum_{i=1}^{n} X_i\):

The likelihood function \(L(\theta; \mathbf{x}) = \theta^n (\prod_{i=1}^{n} x_i)^{\theta-1}\) cannot be expressed as a function of only \(\sum_{i=1}^{n} x_i\) and \(\theta\). For example, if we know \(\sum x_i\), we cannot determine the value of \(\prod x_i\) (e.g., if \(x_1+x_2 = 1\), \(x_1x_2\) could be 0.25 (for \(x_1=x_2=0.5\)) or 0.21 (for \(x_1=0.3, x_2=0.7\))). Therefore, \(\sum_{i=1}^{n} X_i\) is not a sufficient statistic for \(\theta\). Thus, (C) is incorrect.


Step 4: Final Answer:

Based on the analysis, the sufficient statistics for \(\theta\) are \((X_1, \dots, X_n)\), \((X_{(1)}, \dots, X_{(n)})\), and \(\prod_{i=1}^{n} X_i\). Options (A), (B), and (D) are correct.
Quick Tip: The Fisher-Neyman Factorization Theorem is the most direct way to identify sufficient statistics. Always start by writing down the joint pdf (likelihood function) and then try to group terms into a function of the statistic and \(\theta\), and another function of the data only. Remember that any one-to-one function of a sufficient statistic is also sufficient.


Question 39:

A simple linear regression model \(Y_i = \beta_0 + \beta_1 x_i + \epsilon_i\), with \(x_i = (-1)^i\) for \(i=1, 2, \dots, 20\), is fitted. The random error variables \(\epsilon_i\) are uncorrelated with mean 0 and finite variance \(\sigma^2 > 0\). Let \(\hat{\beta}_0\) and \(\hat{\beta}_1\) be the least squares estimators of \(\beta_0\) and \(\beta_1\) respectively. Let \(\hat{Y}_i\) be the fitted value of the \(i^{th}\) response variable \(Y_i\) for \(i=1, \dots, 20\). Which of the following statements is/are correct?

  • (A) \(Cov(\hat{\beta}_0, \hat{\beta}_1) = 0\)
  • (B) \(Var(\hat{\beta}_0) = Var(\hat{\beta}_1)\)
  • (C) \(Var(\hat{\beta}_0) = Cov(\hat{Y}_i, \hat{\beta}_0)\) for all \(i=1, \dots, 20\)
  • (D) \(Var(\hat{\beta}_1) = Cov(\hat{Y}_i, \hat{\beta}_1)\) for all \(i=1, \dots, 20\)
Correct Answer: (A), (B), (C)
View Solution




Step 1: Understanding the Concept:

This question requires an understanding of the properties of least squares estimators in a simple linear regression model. We need to calculate the variance and covariance of the estimators \(\hat{\beta}_0\) and \(\hat{\beta}_1\) and the covariance involving the fitted values \(\hat{Y}_i\).


Step 2: Key Formula or Approach:

The formulas for the variances and covariance of the least squares estimators are: \[ Var(\hat{\beta}_1) = \frac{\sigma^2}{S_{xx}} \] \[ Var(\hat{\beta}_0) = \sigma^2 \left( \frac{1}{n} + \frac{\bar{x}^2}{S_{xx}} \right) \] \[ Cov(\hat{\beta}_0, \hat{\beta}_1) = -\frac{\sigma^2 \bar{x}}{S_{xx}} \]
where \(n\) is the sample size, \(\bar{x} = \frac{1}{n}\sum x_i\), and \(S_{xx} = \sum_{i=1}^n (x_i - \bar{x})^2\).
The fitted value is \(\hat{Y}_i = \hat{\beta}_0 + \hat{\beta}_1 x_i\).


Step 3: Detailed Explanation:

First, let's compute the necessary summary statistics for \(x_i = (-1)^i\) with \(n=20\).
The sequence of \(x_i\) is \(-1, 1, -1, 1, \dots, -1, 1\). There are 10 values of -1 and 10 values of 1. \[ \sum_{i=1}^{20} x_i = 10 \times (-1) + 10 \times (1) = -10 + 10 = 0 \] \[ \bar{x} = \frac{\sum x_i}{n} = \frac{0}{20} = 0 \] \[ \sum_{i=1}^{20} x_i^2 = 10 \times (-1)^2 + 10 \times (1)^2 = 10 \times 1 + 10 \times 1 = 20 \] \[ S_{xx} = \sum (x_i - \bar{x})^2 = \sum x_i^2 - n\bar{x}^2 = 20 - 20(0)^2 = 20 \]
Now, we can evaluate each statement.


(A) \(Cov(\hat{\beta}_0, \hat{\beta}_1) = 0\):

Using the formula for covariance: \[ Cov(\hat{\beta}_0, \hat{\beta}_1) = -\frac{\sigma^2 \bar{x}}{S_{xx}} = -\frac{\sigma^2(0)}{20} = 0 \]
Statement (A) is correct.


(B) \(Var(\hat{\beta}_0) = Var(\hat{\beta}_1)\):

Let's compute the variances: \[ Var(\hat{\beta}_0) = \sigma^2 \left( \frac{1}{n} + \frac{\bar{x}^2}{S_{xx}} \right) = \sigma^2 \left( \frac{1}{20} + \frac{0^2}{20} \right) = \frac{\sigma^2}{20} \] \[ Var(\hat{\beta}_1) = \frac{\sigma^2}{S_{xx}} = \frac{\sigma^2}{20} \]
Since \(Var(\hat{\beta}_0) = Var(\hat{\beta}_1)\), statement (B) is correct.


(C) \(Var(\hat{\beta}_0) = Cov(\hat{Y}_i, \hat{\beta}_0)\) for all \(i\):

Let's compute the covariance \(Cov(\hat{Y}_i, \hat{\beta}_0)\): \[ Cov(\hat{Y}_i, \hat{\beta}_0) = Cov(\hat{\beta}_0 + \hat{\beta}_1 x_i, \hat{\beta}_0) \]
Using the properties of covariance: \[ Cov(\hat{\beta}_0 + \hat{\beta}_1 x_i, \hat{\beta}_0) = Cov(\hat{\beta}_0, \hat{\beta}_0) + x_i Cov(\hat{\beta}_1, \hat{\beta}_0) \] \[ = Var(\hat{\beta}_0) + x_i Cov(\hat{\beta}_0, \hat{\beta}_1) \]
From part (A), we know \(Cov(\hat{\beta}_0, \hat{\beta}_1) = 0\). So, \[ Cov(\hat{Y}_i, \hat{\beta}_0) = Var(\hat{\beta}_0) + x_i(0) = Var(\hat{\beta}_0) \]
This holds for all \(i = 1, \dots, 20\). Statement (C) is correct.


(D) \(Var(\hat{\beta}_1) = Cov(\hat{Y}_i, \hat{\beta}_1)\) for all \(i\):

Let's compute the covariance \(Cov(\hat{Y}_i, \hat{\beta}_1)\): \[ Cov(\hat{Y}_i, \hat{\beta}_1) = Cov(\hat{\beta}_0 + \hat{\beta}_1 x_i, \hat{\beta}_1) \] \[ = Cov(\hat{\beta}_0, \hat{\beta}_1) + x_i Cov(\hat{\beta}_1, \hat{\beta}_1) \] \[ = Cov(\hat{\beta}_0, \hat{\beta}_1) + x_i Var(\hat{\beta}_1) \]
Again, since \(Cov(\hat{\beta}_0, \hat{\beta}_1) = 0\): \[ Cov(\hat{Y}_i, \hat{\beta}_1) = x_i Var(\hat{\beta}_1) \]
The statement claims \(Var(\hat{\beta}_1) = Cov(\hat{Y}_i, \hat{\beta}_1)\), which means \(Var(\hat{\beta}_1) = x_i Var(\hat{\beta}_1)\).
Since \(Var(\hat{\beta}_1) = \sigma^2/20 > 0\), this would imply \(x_i = 1\). However, \(x_i\) can be -1 (for odd \(i\)). Thus, the statement is not true for all \(i\). Statement (D) is incorrect.


Step 4: Final Answer:

Statements (A), (B), and (C) are correct.
Quick Tip: In simple linear regression, if the predictor variable \(x\) is centered (i.e., \(\bar{x} = 0\)), the estimators for the intercept (\(\hat{\beta}_0\)) and slope (\(\hat{\beta}_1\)) become uncorrelated. This simplifies many calculations and is a useful property to check for at the start of a problem.


Question 40:

Let \(Y_1, Y_2, \dots, Y_n\) be i.i.d. discrete random variables from a population with probability mass function

where \(0 < \theta < 1\) is an unknown parameter. Assume the convention \(0^0=1\). If \(\hat{\theta}\) is the method of moments estimator of \(\theta\), then which of the following statements is/are correct?

  • (A) \(\hat{\theta}\) is also the maximum likelihood estimator of \(\theta\)
  • (B) \(\hat{\theta}\) is an unbiased estimator of \(\theta\)
  • (C) \(\hat{\theta}\) is a consistent estimator of \(\theta\)
  • (D) \(1/\hat{\theta}\) is an unbiased estimator of \(1/\theta\)
Correct Answer: (A), (C), (D)
View Solution




Step 1: Understanding the Concept:

The question concerns a Geometric distribution \(Y \sim Geo(\theta)\) (number of failures before the first success). We need to find the method of moments estimator (MME) and the maximum likelihood estimator (MLE) for the parameter \(\theta\) and then check its properties like unbiasedness and consistency.


Step 2: Detailed Explanation:

The given distribution is a Geometric distribution with parameter \(\theta\). The expected value is \(E[Y] = \frac{1-\theta}{\theta}\).


Finding the Method of Moments Estimator (MME):

The method of moments equates the first population moment \(E[Y]\) to the first sample moment \(\bar{Y} = \frac{1}{n}\sum Y_i\). \[ E[Y] = \bar{Y} \] \[ \frac{1-\theta}{\theta} = \bar{Y} \]
Solving for \(\theta\): \[ 1-\theta = \bar{Y}\theta \] \[ 1 = \theta(1+\bar{Y}) \] \[ \hat{\theta}_{MME} = \frac{1}{1+\bar{Y}} \]

(A) Checking if \(\hat{\theta}_{MME}\) is also the MLE:

The likelihood function is: \[ L(\theta) = \prod_{i=1}^{n} P(Y_i=y_i; \theta) = \prod_{i=1}^{n} \theta(1-\theta)^{y_i} = \theta^n (1-\theta)^{\sum y_i} \]
The log-likelihood function is: \[ \ln L(\theta) = n \ln(\theta) + \left(\sum y_i\right) \ln(1-\theta) \]
To find the MLE, we differentiate with respect to \(\theta\) and set it to zero: \[ \frac{d}{d\theta} \ln L(\theta) = \frac{n}{\theta} - \frac{\sum y_i}{1-\theta} = 0 \] \[ \frac{n}{\hat{\theta}_{MLE}} = \frac{\sum y_i}{1-\hat{\theta}_{MLE}} = \frac{n\bar{Y}}{1-\hat{\theta}_{MLE}} \] \[ n(1-\hat{\theta}_{MLE}) = n\bar{Y}\hat{\theta}_{MLE} \] \[ 1 - \hat{\theta}_{MLE} = \bar{Y}\hat{\theta}_{MLE} \] \[ 1 = \hat{\theta}_{MLE}(1+\bar{Y}) \] \[ \hat{\theta}_{MLE} = \frac{1}{1+\bar{Y}} \]
Since \(\hat{\theta}_{MME} = \hat{\theta}_{MLE}\), statement (A) is correct. Let's denote this common estimator by \(\hat{\theta}\).


(B) Checking if \(\hat{\theta}\) is unbiased:

We need to check if \(E[\hat{\theta}] = \theta\). \[ E[\hat{\theta}] = E\left[\frac{1}{1+\bar{Y}}\right] \]
The function \(g(x) = \frac{1}{1+x}\) is convex for \(x > -1\). Since \(\bar{Y} \ge 0\), we can apply Jensen's inequality, which states that for a convex function \(g\), \(E[g(X)] \ge g(E[X])\). \[ E[\hat{\theta}] \ge \frac{1}{1+E[\bar{Y}]} \]
We know \(E[\bar{Y}] = E[Y] = \frac{1-\theta}{\theta}\). \[ E[\hat{\theta}] \ge \frac{1}{1+\frac{1-\theta}{\theta}} = \frac{1}{\frac{\theta+1-\theta}{\theta}} = \frac{1}{1/\theta} = \theta \]
Since the inequality is strict for a non-constant random variable \(\bar{Y}\), we have \(E[\hat{\theta}] > \theta\). Therefore, \(\hat{\theta}\) is a biased estimator. Statement (B) is incorrect.


(C) Checking if \(\hat{\theta}\) is consistent:

An estimator is consistent if it converges in probability to the true parameter value as \(n \to \infty\). By the Weak Law of Large Numbers, the sample mean converges in probability to the population mean: \[ \bar{Y} \xrightarrow{p} E[Y] = \frac{1-\theta}{\theta} \] \(\hat{\theta} = g(\bar{Y}) = \frac{1}{1+\bar{Y}}\). Since \(g(x)=\frac{1}{1+x}\) is a continuous function at \(x = E[Y]\), by the Continuous Mapping Theorem: \[ \hat{\theta} = \frac{1}{1+\bar{Y}} \xrightarrow{p} \frac{1}{1+E[Y]} = \frac{1}{1+\frac{1-\theta}{\theta}} = \theta \]
Thus, \(\hat{\theta}\) is a consistent estimator of \(\theta\). Statement (C) is correct.


(D) Checking if \(1/\hat{\theta}\) is an unbiased estimator of \(1/\theta\):

We have \(1/\hat{\theta} = 1 / \left(\frac{1}{1+\bar{Y}}\right) = 1+\bar{Y}\).
Let's find its expected value: \[ E\left[\frac{1}{\hat{\theta}}\right] = E[1+\bar{Y}] = 1 + E[\bar{Y}] = 1 + E[Y] = 1 + \frac{1-\theta}{\theta} \]
Simplifying the expression: \[ 1 + \frac{1-\theta}{\theta} = \frac{\theta}{\theta} + \frac{1-\theta}{\theta} = \frac{\theta+1-\theta}{\theta} = \frac{1}{\theta} \]
So, \(E[1/\hat{\theta}] = 1/\theta\). This means \(1/\hat{\theta}\) is an unbiased estimator of \(1/\theta\). Statement (D) is correct.


Step 4: Final Answer:

Statements (A), (C), and (D) are correct.
Quick Tip: For distributions in the exponential family (like the Geometric distribution), the MME and MLE often coincide. To check for consistency, rely on the Law of Large Numbers and the Continuous Mapping Theorem. For unbiasedness, calculating the expectation directly is required; Jensen's inequality is a powerful tool to prove bias without computing the exact expectation.


Question 41:

Let \(A\) be a \(7 \times 7\) real matrix with \(rank(A) = 1\). Suppose the trace of \(A^2\) is 2025. Let the characteristic polynomial of \(A\) be written as \(\sum_{n=0}^{7} a_n x^n\). Then \(\sum_{n=0}^{7} |a_n|\) is ___________.

(answer in integer)

Correct Answer: 46
View Solution




Step 1: Understanding the Concept:

The properties of a rank-1 matrix are key to solving this problem. A matrix of rank 1 has at most one non-zero eigenvalue. The trace of a matrix is the sum of its eigenvalues, and the trace of its square is the sum of the squares of its eigenvalues.


Step 2: Key Formula or Approach:

1. For an \(n \times n\) matrix \(A\) with \(rank(A)=1\), it has one non-zero eigenvalue, \(\lambda\), and \(n-1\) eigenvalues equal to 0.

2. The non-zero eigenvalue is equal to the trace of the matrix, i.e., \(\lambda = tr(A)\).

3. The eigenvalues of \(A^2\) are the squares of the eigenvalues of \(A\).

4. The characteristic polynomial is given by \(p(x) = \det(A - xI)\).


Step 3: Detailed Explanation:

Given that \(A\) is a \(7 \times 7\) matrix with \(rank(A) = 1\), it has one non-zero eigenvalue, let's call it \(\lambda\), and the other six eigenvalues are 0.

The eigenvalues of \(A\) are \(\{\lambda, 0, 0, 0, 0, 0, 0\}\).

The eigenvalues of \(A^2\) are the squares of the eigenvalues of \(A\), which are \(\{\lambda^2, 0^2, 0^2, 0^2, 0^2, 0^2, 0^2\} = \{\lambda^2, 0, 0, 0, 0, 0, 0\}\).

The trace of \(A^2\) is the sum of its eigenvalues.
\[ tr(A^2) = \lambda^2 + 0 + 0 + 0 + 0 + 0 + 0 = \lambda^2 \]
We are given that \(tr(A^2) = 2025\).
\[ \lambda^2 = 2025 \implies \lambda = \pm\sqrt{2025} = \pm 45 \]
The characteristic polynomial of \(A\) is usually defined as \(\det(xI - A)\). However, the problem states it as \(\sum a_n x^n\), which is typically the form for \(\det(A - xI)\). Let's work with this form.

The characteristic polynomial is the product of \((\lambda_i - x)\) for each eigenvalue \(\lambda_i\).
\[ p(x) = \det(A - xI) = (\lambda - x)(0 - x)(0 - x)(0 - x)(0 - x)(0 - x)(0 - x) \] \[ p(x) = (\lambda - x)(-x)^6 = (\lambda - x)x^6 = \lambda x^6 - x^7 \]
This polynomial is written in the form \(\sum_{n=0}^{7} a_n x^n = a_0 + a_1 x + \dots + a_7 x^7\).

Comparing the terms of our polynomial \(p(x) = \lambda x^6 - x^7\) with the general form:
\(a_7 = -1\)
\(a_6 = \lambda\)

All other coefficients \(a_n\) for \(n \in \{0, 1, 2, 3, 4, 5\}\) are zero.

The question asks for the value of \(\sum_{n=0}^{7} |a_n|\).
\[ \sum_{n=0}^{7} |a_n| = |a_0| + |a_1| + \dots + |a_5| + |a_6| + |a_7| \] \[ \sum_{n=0}^{7} |a_n| = 0 + 0 + \dots + 0 + |\lambda| + |-1| \] \[ \sum_{n=0}^{7} |a_n| = |\lambda| + 1 \]
Since \(\lambda = \pm 45\), we have \(|\lambda| = 45\).
\[ \sum_{n=0}^{7} |a_n| = 45 + 1 = 46 \]

Step 4: Final Answer:

The value of \(\sum_{n=0}^{7} |a_n|\) is 46.
Quick Tip: For a rank-1 matrix, remember that it has only one non-zero eigenvalue, which is equal to its trace. The characteristic polynomial is then simple to compute. The trace of any power of the matrix, \(tr(A^k)\), is simply the sum of the k-th powers of its eigenvalues.


Question 42:

The radius of convergence of the power series \(\sum_{n=1}^{\infty} \frac{n!}{n^n} x^n\) is equal to ___________.

(round off to 2 decimal places)

Correct Answer: 2.72
View Solution




Step 1: Understanding the Concept:

The radius of convergence of a power series determines the interval on which the series converges. A common method to find the radius of convergence is the Ratio Test.


Step 2: Key Formula or Approach:

For a power series \(\sum_{n=1}^{\infty} c_n x^n\), the radius of convergence \(R\) can be calculated using the Ratio Test: \[ R = \lim_{n \to \infty} \left| \frac{c_n}{c_{n+1}} \right| \]
In this problem, the coefficient is \(c_n = \frac{n!}{n^n}\).


Step 3: Detailed Explanation:

We are given the power series \(\sum_{n=1}^{\infty} \frac{n!}{n^n} x^n\).

The coefficient is \(c_n = \frac{n!}{n^n}\).

The next coefficient in the series is \(c_{n+1} = \frac{(n+1)!}{(n+1)^{n+1}}\).

Now, we compute the ratio \(\frac{c_{n+1}}{c_n}\): \[ \frac{c_{n+1}}{c_n} = \frac{(n+1)! / (n+1)^{n+1}}{n! / n^n} = \frac{(n+1)!}{(n+1)^{n+1}} \cdot \frac{n^n}{n!} \]
We can simplify the factorial term: \((n+1)! = (n+1) \cdot n!\).
\[ \frac{c_{n+1}}{c_n} = \frac{(n+1) \cdot n!}{(n+1)^{n+1}} \cdot \frac{n^n}{n!} = \frac{n+1}{(n+1)^{n+1}} \cdot n^n = \frac{n^n}{(n+1)^n} \]
This can be rewritten as: \[ \frac{c_{n+1}}{c_n} = \left(\frac{n}{n+1}\right)^n = \left(\frac{1}{\frac{n+1}{n}}\right)^n = \left(\frac{1}{1 + \frac{1}{n}}\right)^n = \frac{1}{\left(1 + \frac{1}{n}\right)^n} \]
According to the Ratio Test for power series, the series converges if \(\lim_{n \to \infty} \left| \frac{c_{n+1} x^{n+1}}{c_n x^n} \right| < 1\), which means \(|x| \lim_{n \to \infty} \left| \frac{c_{n+1}}{c_n} \right| < 1\).

Let \(L = \lim_{n \to \infty} \left| \frac{c_{n+1}}{c_n} \right|\). The radius of convergence is \(R = 1/L\).
\[ L = \lim_{n \to \infty} \frac{1}{\left(1 + \frac{1}{n}\right)^n} \]
We use the well-known limit: \(\lim_{n \to \infty} \left(1 + \frac{1}{n}\right)^n = e\).
\[ L = \frac{1}{e} \]
The radius of convergence is \(R = \frac{1}{L} = \frac{1}{1/e} = e\).

The value of \(e\) is approximately 2.71828...


Step 4: Final Answer:

Rounding the value of \(e\) to two decimal places, we get 2.72.
Quick Tip: The Ratio Test is extremely useful for series involving factorials and powers of \(n\). Memorizing the fundamental limit \(\lim_{n \to \infty} (1 + x/n)^n = e^x\) is crucial for solving many convergence problems quickly.


Question 43:

Let \(f(x) = x \sin\left(\frac{\pi}{2x}\right), x > 0\). Then \[ \lim_{h \to 0} \frac{1}{\pi^2 h^2} [3f(1) - 2f(1+h) - f(1-2h)] \]
is equal to ___________.

(round off to 2 decimal places)

Correct Answer: 0.75
View Solution




Step 1: Understanding the Concept:

The given limit is in the indeterminate form \(\frac{0}{0}\). This suggests the use of L'Hôpital's Rule or Taylor series expansion. The structure of the limit is related to the second derivative of the function \(f(x)\) at \(x=1\).


Step 2: Key Formula or Approach:

We will apply L'Hôpital's Rule twice, as the limit remains in the \(\frac{0}{0}\) form after the first application. Let the expression inside the limit be \(\frac{N(h)}{D(h)}\), where \(N(h) = 3f(1) - 2f(1+h) - f(1-2h)\) and \(D(h) = \pi^2 h^2\).

We need to find the first and second derivatives of \(f(x)\).


Step 3: Detailed Explanation:

First, let's find the derivatives of \(f(x) = x \sin(\frac{\pi}{2x})\).

Using the product rule and chain rule: \[ f'(x) = (1) \sin\left(\frac{\pi}{2x}\right) + x \cos\left(\frac{\pi}{2x}\right) \cdot \left(-\frac{\pi}{2x^2}\right) = \sin\left(\frac{\pi}{2x}\right) - \frac{\pi}{2x} \cos\left(\frac{\pi}{2x}\right) \]
Now, let's find the second derivative, \(f''(x)\): \[ f''(x) = \frac{d}{dx} \left( \sin\left(\frac{\pi}{2x}\right) \right) - \frac{d}{dx} \left( \frac{\pi}{2x} \cos\left(\frac{\pi}{2x}\right) \right) \] \[ \frac{d}{dx} \left( \sin\left(\frac{\pi}{2x}\right) \right) = \cos\left(\frac{\pi}{2x}\right) \cdot \left(-\frac{\pi}{2x^2}\right) = -\frac{\pi}{2x^2} \cos\left(\frac{\pi}{2x}\right) \] \[ \frac{d}{dx} \left( \frac{\pi}{2x} \cos\left(\frac{\pi}{2x}\right) \right) = \left(-\frac{\pi}{2x^2}\right)\cos\left(\frac{\pi}{2x}\right) + \frac{\pi}{2x}\left(-\sin\left(\frac{\pi}{2x}\right)\right)\left(-\frac{\pi}{2x^2}\right) = -\frac{\pi}{2x^2}\cos\left(\frac{\pi}{2x}\right) + \frac{\pi^2}{4x^3}\sin\left(\frac{\pi}{2x}\right) \] \[ f''(x) = -\frac{\pi}{2x^2} \cos\left(\frac{\pi}{2x}\right) - \left(-\frac{\pi}{2x^2}\cos\left(\frac{\pi}{2x}\right) + \frac{\pi^2}{4x^3}\sin\left(\frac{\pi}{2x}\right)\right) \] \[ f''(x) = -\frac{\pi}{2x^2} \cos\left(\frac{\pi}{2x}\right) + \frac{\pi}{2x^2}\cos\left(\frac{\pi}{2x}\right) - \frac{\pi^2}{4x^3}\sin\left(\frac{\pi}{2x}\right) = -\frac{\pi^2}{4x^3}\sin\left(\frac{\pi}{2x}\right) \]
Let's evaluate \(f''(1)\): \[ f''(1) = -\frac{\pi^2}{4(1)^3}\sin\left(\frac{\pi}{2(1)}\right) = -\frac{\pi^2}{4}\sin\left(\frac{\pi}{2}\right) = -\frac{\pi^2}{4}(1) = -\frac{\pi^2}{4} \]
Now, let's evaluate the limit \(L = \lim_{h \to 0} \frac{3f(1) - 2f(1+h) - f(1-2h)}{\pi^2 h^2}\).
As \(h \to 0\), the numerator becomes \(3f(1) - 2f(1) - f(1) = 0\), and the denominator is 0. Applying L'Hôpital's Rule (differentiating w.r.t. \(h\)): \[ L = \lim_{h \to 0} \frac{-2f'(1+h) - f'(1-2h)(-2)}{2\pi^2 h} = \lim_{h \to 0} \frac{-2f'(1+h) + 2f'(1-2h)}{2\pi^2 h} \]
This is still \(\frac{0}{0}\). Applying L'Hôpital's Rule again: \[ L = \lim_{h \to 0} \frac{-2f''(1+h)(1) + 2f''(1-2h)(-2)}{2\pi^2} = \frac{-2f''(1) - 4f''(1)}{2\pi^2} = \frac{-6f''(1)}{2\pi^2} = \frac{-3f''(1)}{\pi^2} \]
Substitute the value of \(f''(1) = -\frac{\pi^2}{4}\): \[ L = \frac{-3(-\pi^2/4)}{\pi^2} = \frac{3\pi^2/4}{\pi^2} = \frac{3}{4} = 0.75 \]

Step 4: Final Answer:

The value of the limit is 0.75.
Quick Tip: For limits of the form \(\frac{0}{0}\) involving function evaluations, L'Hôpital's Rule is a powerful tool. Alternatively, using Taylor series expansions for \(f(x+h)\) around \(x\) can simplify the expression and reveal its relation to derivatives, often providing a quicker path to the solution.


Question 44:

Let \(\{X_i\}_{i \ge 1}\) be a sequence of i.i.d. random variables with \(E(X_i) = \mu\) and \(Var(X_i) = \sigma^2 < \infty\). Suppose \(c\) is a constant that does not depend on \(n\) such that \[ \frac{c}{n} \sum_{i=1}^{n} (X_{2i} - X_{2i-1})^2 \]
is a consistent estimator of \(\sigma^2\). Then \(c\) is equal to ___________.

(round off to 2 decimal places)

Correct Answer: 0.50
View Solution




Step 1: Understanding the Concept:

An estimator \(T_n\) for a parameter \(\theta\) is consistent if it converges in probability to \(\theta\) as the sample size \(n\) approaches infinity. A sufficient condition for consistency is that the estimator is asymptotically unbiased (\( \lim_{n \to \infty} E(T_n) = \theta \)) and its variance approaches zero (\( \lim_{n \to \infty} Var(T_n) = 0 \)). A simple way to ensure this for many estimators is to make them unbiased, i.e., \(E(T_n) = \theta\).


Step 2: Key Formula or Approach:

Let the estimator be \(T_n = \frac{c}{n} \sum_{i=1}^{n} (X_{2i} - X_{2i-1})^2\). We will find the value of \(c\) that makes \(T_n\) an unbiased estimator of \(\sigma^2\). That is, we will solve the equation \(E(T_n) = \sigma^2\) for \(c\).

We will use the properties of expectation and variance:
1. \(E(A-B) = E(A) - E(B)\)

2. \(Var(A-B) = Var(A) + Var(B)\) if A and B are independent.

3. \(Var(Y) = E(Y^2) - [E(Y)]^2 \implies E(Y^2) = Var(Y) + [E(Y)]^2\).


Step 3: Detailed Explanation:

Let's find the expected value of the estimator \(T_n\). \[ E(T_n) = E\left[\frac{c}{n} \sum_{i=1}^{n} (X_{2i} - X_{2i-1})^2\right] \]
By the linearity of expectation, we can move the constant and the summation outside: \[ E(T_n) = \frac{c}{n} \sum_{i=1}^{n} E\left[(X_{2i} - X_{2i-1})^2\right] \]
Let \(Y_i = X_{2i} - X_{2i-1}\). We need to find \(E(Y_i^2)\).

First, find the expectation of \(Y_i\): \[ E(Y_i) = E(X_{2i} - X_{2i-1}) = E(X_{2i}) - E(X_{2i-1}) = \mu - \mu = 0 \]
Next, find the variance of \(Y_i\). Since the \(X_i\) are i.i.d., \(X_{2i}\) and \(X_{2i-1}\) are independent. \[ Var(Y_i) = Var(X_{2i} - X_{2i-1}) = Var(X_{2i}) + Var(X_{2i-1}) = \sigma^2 + \sigma^2 = 2\sigma^2 \]
Now, we can find \(E(Y_i^2)\) using the variance formula: \[ E(Y_i^2) = Var(Y_i) + [E(Y_i)]^2 = 2\sigma^2 + (0)^2 = 2\sigma^2 \]
So, \(E[(X_{2i} - X_{2i-1})^2] = 2\sigma^2\). This value is the same for all \(i\).

Substitute this back into the expression for \(E(T_n)\): \[ E(T_n) = \frac{c}{n} \sum_{i=1}^{n} (2\sigma^2) = \frac{c}{n} (n \cdot 2\sigma^2) = 2c\sigma^2 \]
For \(T_n\) to be an unbiased (and hence consistent, provided variance goes to 0) estimator of \(\sigma^2\), its expected value must be equal to \(\sigma^2\). \[ E(T_n) = \sigma^2 \] \[ 2c\sigma^2 = \sigma^2 \]
Assuming \(\sigma^2 > 0\), we can divide by \(\sigma^2\): \[ 2c = 1 \implies c = \frac{1}{2} = 0.5 \]
This choice of c makes the estimator unbiased. By the Law of Large Numbers, this sample mean of i.i.d. terms \((X_{2i}-X_{2i-1})^2\) will converge to its expected value, ensuring consistency.

Step 4: Final Answer:

The value of \(c\) is 0.50.
Quick Tip: When asked to find a constant that makes an estimator consistent, a good first step is to find the constant that makes it unbiased. This often provides the correct answer. Remember the fundamental relationship \(E[Y^2] = Var(Y) + [E(Y)]^2\), which is frequently used to find the expected value of a squared term.


Question 45:

Let \(X_1, X_2, \dots, X_{10}\) be i.i.d. U\((0, \theta)\) random variables, where \(\theta > 0\) is unknown. For testing the null hypothesis \(H_0: \theta = 1\) against the alternative hypothesis \(H_1: \theta = 0.9\), consider a test that rejects \(H_0\) if \[ X_{(10)} = \max\{X_1, X_2, \dots, X_{10}\} < 0.8. \]
Then the probability of type I error of the test is equal to ___________.

(round off to 2 decimal places)

Correct Answer: 0.11
View Solution




Step 1: Understanding the Concept:

A Type I error occurs when we reject the null hypothesis \(H_0\) when it is actually true. The probability of a Type I error is denoted by \(\alpha\). In this problem, we need to calculate the probability of the rejection event, assuming that \(H_0\) is true.


Step 2: Key Formula or Approach:

1. The probability of a Type I error is \(\alpha = P(Reject H_0 | H_0 is true)\).

2. The rejection region is given by \(X_{(10)} < 0.8\).

3. Under the null hypothesis \(H_0\), the parameter \(\theta = 1\), which means \(X_i\) are i.i.d. random variables from a Uniform(0, 1) distribution.

4. For \(n\) i.i.d. random variables \(X_1, \dots, X_n\) with CDF \(F_X(x)\), the CDF of the maximum order statistic \(X_{(n)}\) is given by \(F_{X_{(n)}}(x) = [F_X(x)]^n\).


Step 3: Detailed Explanation:

The probability of a Type I error is the probability of rejecting \(H_0\) given that \(H_0\) is true.
\[ \alpha = P(X_{(10)} < 0.8 | H_0 is true) \]
Under \(H_0\), we have \(\theta = 1\). Thus, \(X_1, X_2, \dots, X_{10}\) are i.i.d. U(0, 1).

The cumulative distribution function (CDF) for a single \(X_i \sim U(0,1)\) is \(F_X(x) = x\) for \(0 < x < 1\).

The CDF of the maximum order statistic, \(X_{(10)}\), is: \[ F_{X_{(10)}}(x) = P(X_{(10)} \le x) = P(\max\{X_1, \dots, X_{10}\} \le x) \]
This is equivalent to all \(X_i\) being less than or equal to \(x\): \[ P(X_1 \le x, X_2 \le x, \dots, X_{10} \le x) \]
Since the variables are i.i.d., this probability is the product of individual probabilities: \[ F_{X_{(10)}}(x) = [P(X_1 \le x)]^{10} = [F_X(x)]^{10} = x^{10} \quad for 0 < x < 1 \]
We need to find the probability that \(X_{(10)} < 0.8\). This is the value of the CDF of \(X_{(10)}\) at \(x=0.8\).
\[ \alpha = F_{X_{(10)}}(0.8) = (0.8)^{10} \]
Now we calculate the value: \[ (0.8)^{10} = \left(\frac{4}{5}\right)^{10} = \frac{4^{10}}{5^{10}} = \frac{1,048,576}{9,765,625} \approx 0.107374 \]

Step 4: Final Answer:

Rounding the result to 2 decimal places, we get 0.11.
Quick Tip: The probability of a Type I error, \(\alpha\), is the size of the critical region under the null hypothesis. For order statistics problems, remember that \(P(\max(X_i) \le x)\) is the probability that *all* \(X_i\) are \(\le x\), and \(P(\min(X_i) > x)\) is the probability that *all* \(X_i\) are \(> x\). These are often easier to work with than the PDFs.


Question 46:

Let \((X_1, Y_1), (X_2, Y_2), \dots, (X_{100}, Y_{100})\) be i.i.d. discrete random vectors each having joint probability mass function \[ P(X=x, Y=y) = \frac{e^{-(1+x)\lambda}((1+x)\lambda)^y}{y!} p^x (1-p)^{1-x}, \quad x \in \{0, 1\}, y \in \mathbb{N} \cup \{0\}, \]
where \(\lambda > 0\) and \(0 < p < 1\) are unknown parameters. If the observed values of \(\sum_{i=1}^{100} X_i\) and \(\sum_{i=1}^{100} Y_i\) are 54 and 521 respectively, the maximum likelihood estimate of \(\lambda\) is equal to ___________.

(round off to 2 decimal places)

Correct Answer: 3.38
View Solution




Step 1: Understanding the Concept:

Maximum Likelihood Estimation (MLE) is a method for estimating the parameters of a statistical model. The principle is to find the parameter values that maximize the likelihood function, which is the joint probability of observing the given data.


Step 2: Key Formula or Approach:

1. Write down the likelihood function \(L(\lambda, p)\), which is the product of the PMFs for each observation.

2. Take the natural logarithm of the likelihood function, \(\ln L\), to simplify the product into a sum.

3. To find the MLE for \(\lambda\), take the partial derivative of \(\ln L\) with respect to \(\lambda\) and set it to zero.

4. Solve the resulting equation for \(\lambda\).


Step 3: Detailed Explanation:

The likelihood function for \(n=100\) observations \((x_i, y_i)\) is: \[ L(\lambda, p) = \prod_{i=1}^{100} P(X_i=x_i, Y_i=y_i) = \prod_{i=1}^{100} \left[ \frac{e^{-(1+x_i)\lambda}((1+x_i)\lambda)^{y_i}}{y_i!} p^{x_i} (1-p)^{1-x_i} \right] \]
The log-likelihood function is: \[ \ln L = \sum_{i=1}^{100} \ln \left[ \frac{e^{-(1+x_i)\lambda}((1+x_i)\lambda)^{y_i}}{y_i!} p^{x_i} (1-p)^{1-x_i} \right] \] \[ \ln L = \sum_{i=1}^{100} \left[ -(1+x_i)\lambda + y_i \ln((1+x_i)\lambda) - \ln(y_i!) + x_i \ln p + (1-x_i)\ln(1-p) \right] \]
We can separate the terms involving \(\lambda\) and \(p\). To find the MLE of \(\lambda\), we only need the terms containing \(\lambda\). \[ \ln L = -\lambda \sum_{i=1}^{100}(1+x_i) + \sum_{i=1}^{100} y_i \ln(1+x_i) + \ln(\lambda) \sum_{i=1}^{100} y_i + (terms not involving \lambda) \] \[ \ln L = -\lambda \left(100 + \sum x_i\right) + \sum y_i \ln(1+x_i) + \ln(\lambda) \sum y_i + \dots \]
Now, we differentiate with respect to \(\lambda\) and set the derivative to zero: \[ \frac{\partial \ln L}{\partial \lambda} = -\left(100 + \sum x_i\right) + \frac{1}{\lambda} \sum y_i = 0 \]
Solving for the MLE \(\hat{\lambda}\): \[ \frac{1}{\hat{\lambda}} \sum y_i = 100 + \sum x_i \] \[ \hat{\lambda} = \frac{\sum_{i=1}^{100} y_i}{100 + \sum_{i=1}^{100} x_i} \]
We are given \(\sum_{i=1}^{100} X_i = 54\) and \(\sum_{i=1}^{100} Y_i = 521\). \[ \hat{\lambda} = \frac{521}{100 + 54} = \frac{521}{154} \] \[ \hat{\lambda} \approx 3.383116... \]

Step 4: Final Answer:

Rounding the result to 2 decimal places, we get 3.38.
Quick Tip: Notice that the joint PMF can be factored as \(P(Y=y|X=x)P(X=x)\), where \(X \sim Bernoulli(p)\) and \(Y|X=x \sim Poisson((1+x)\lambda)\). The log-likelihood function separates into terms involving only \(\lambda\) and terms involving only \(p\). This means you can find the MLE for each parameter separately by maximizing the corresponding part of the log-likelihood.


Question 47:

Let \(X\) and \(Y\) be i.i.d. random variables with probability density function

Suppose \(Z = \min\{X, Y\}\), then \(E(Z)\) is equal to ___________.

(answer in integer)

Correct Answer: 2
View Solution




Step 1: Understanding the Concept:

We need to find the expected value of the minimum of two i.i.d. random variables. A common strategy is to first find the probability distribution (either CDF or PDF) of the minimum, and then use the definition of expected value.


Step 2: Key Formula or Approach:

1. Find the survival function \(S_X(x) = P(X > x)\) for a single random variable.

2. Find the survival function of \(Z\), \(S_Z(z) = P(Z > z)\). Since \(Z = \min\{X, Y\}\), \(Z > z\) if and only if both \(X > z\) and \(Y > z\).

3. Use the survival function to find the PDF of \(Z\), \(f_Z(z) = -S'_Z(z)\).

4. Calculate the expected value \(E(Z) = \int z f_Z(z) dz\) over the support of \(Z\).


Step 3: Detailed Explanation:

First, we find the survival function of \(X\). For \(x \ge 1\): \[ S_X(x) = P(X > x) = \int_x^{\infty} f(t) dt = \int_x^{\infty} \frac{1}{t^2} dt = \left[-\frac{1}{t}\right]_x^{\infty} = 0 - \left(-\frac{1}{x}\right) = \frac{1}{x} \]
Next, we find the survival function of \(Z = \min\{X, Y\}\). Since \(X, Y \ge 1\), we have \(Z \ge 1\). For \(z \ge 1\): \[ S_Z(z) = P(Z > z) = P(\min\{X, Y\} > z) = P(X > z and Y > z) \]
Since \(X\) and \(Y\) are i.i.d.: \[ S_Z(z) = P(X > z) \cdot P(Y > z) = S_X(z) \cdot S_Y(z) = \left(\frac{1}{z}\right) \cdot \left(\frac{1}{z}\right) = \frac{1}{z^2} \]
Now, we find the PDF of \(Z\) from its survival function: \[ f_Z(z) = -\frac{d}{dz} S_Z(z) = -\frac{d}{dz} \left(z^{-2}\right) = -(-2z^{-3}) = \frac{2}{z^3} \quad for z \ge 1 \]
Finally, we calculate the expected value of \(Z\): \[ E(Z) = \int_1^{\infty} z \cdot f_Z(z) dz = \int_1^{\infty} z \cdot \left(\frac{2}{z^3}\right) dz = \int_1^{\infty} \frac{2}{z^2} dz \] \[ E(Z) = 2 \left[-\frac{1}{z}\right]_1^{\infty} = 2 \left( 0 - \left(-\frac{1}{1}\right) \right) = 2(1) = 2 \]

Step 4: Final Answer:

The expected value of \(Z\) is 2.
Quick Tip: An alternative method to compute the expectation of a non-negative random variable \(Z\) is using its survival function directly: \(E(Z) = \int_0^{\infty} S_Z(z) dz\). For a variable with support \([a, \infty)\), the formula is \(E(Z) = a + \int_a^\infty S_Z(z) dz\). In this case, \(a=1\), so \(E(Z) = 1 + \int_1^\infty (1/z^2) dz = 1+1=2\). This can sometimes be faster than finding the PDF first.


Question 48:

The joint probability density function of the random vector \((X, Y, Z)\) is given by

Then the value of \(P(X > 5Y)\) is equal to ___________.

(round off to 2 decimal places)

Correct Answer: 0.20
View Solution




Step 1: Understanding the Concept:

We are given a joint PDF for three variables \((X, Y, Z)\) and asked to find the probability of an event involving only \(X\) and \(Y\). This requires integrating the joint PDF over the specified region. A crucial first step is to check if the given function is a valid PDF (i.e., integrates to 1). If not, we must normalize it.


Step 2: Key Formula or Approach:

1. Check for normalization: Integrate \(f(x, y, z)\) over its entire support to find the total probability. Let this be \(C\).

2. The correct PDF is \(f_{true}(x, y, z) = \frac{1}{C} f(x, y, z)\).

3. Find the marginal PDF of \((X, Y)\) by integrating the true PDF with respect to \(z\).

4. Calculate \(P(X > 5Y)\) by integrating the marginal PDF of \((X, Y)\) over the region defined by \(0 < y < x < 1\) and \(x > 5y\).


Step 3: Detailed Explanation:


The condition \(x > 5y\) implies \(y < x/5\). Since we already have \(y < x\), the stricter condition is \(y < x/5\).

So the region is \(0 < y < x/5\) and \(0 < x < 1\).

We set up the integral: \[ P(X > 5Y) = \iint_{x>5y, 0 < y < x < 1} f_{X,Y}(x,y) dA = \int_0^1 \int_0^{x/5} 2 \, dy \, dx \]
First, we integrate with respect to \(y\):
\[ \int_0^{x/5} 2 \, dy = [2y]_0^{x/5} = 2\left(\frac{x}{5}\right) - 0 = \frac{2x}{5} \]
Now, we integrate the result with respect to \(x\):
\[ \int_0^1 \frac{2x}{5} \, dx = \frac{2}{5} \left[\frac{x^2}{2}\right]_0^1 = \frac{2}{5} \left(\frac{1^2}{2} - 0\right) = \frac{2}{5} \cdot \frac{1}{2} = \frac{1}{5} = 0.2 \]

Step 4: Final Answer:

The value of the probability is 0.20.
Quick Tip: If a calculation based on a given PDF in an exam question leads to a result that is very different from the options or the answer key, double-check the normalization of the PDF. Many questions provide a function proportional to the PDF, and you are expected to find the normalization constant yourself. If the result is still inconsistent, consider plausible typos, such as assuming a uniform distribution if the geometry is simple.


Question 49:

Suppose \( U_1 \) and \( U_2 \) are i.i.d. U(0,1) random variables. Further, let X be a Bin(2, 0.5) random variable that is independent of \( (U_1, U_2) \). Then \[ 36 \mathbb{P}(U_1 + U_2 > X) \]
is equal to ________________
(answer in integer)

Correct Answer: 18
View Solution




Step 1: Understanding the Concept:

We need to calculate the probability \( \mathbb{P}(U_1 + U_2 > X) \). Since \( X \) is a discrete random variable and independent of \( U_1 \) and \( U_2 \), we can use the law of total probability by conditioning on the possible values of \( X \).


Step 2: Key Formula or Approach:

The law of total probability states: \[ \mathbb{P}(A) = \sum_{i} \mathbb{P}(A | B_i) \mathbb{P}(B_i) \]
In this problem, event A is \( U_1 + U_2 > X \) and events \( B_i \) are \( X=k \) for the possible values of \( k \).

The probability mass function for a binomial distribution \( Bin(n, p) \) is \( \mathbb{P}(X=k) = \binom{n}{k} p^k (1-p)^{n-k} \).

The probability \( \mathbb{P}(U_1 + U_2 \leq s) \) for \( U_1, U_2 \sim U(0,1) \) can be found by considering the area in the unit square.


Step 3: Detailed Explanation:

First, let's find the probability distribution of \( X \sim Bin(2, 0.5) \). The possible values for \( X \) are 0, 1, and 2.
\[ \mathbb{P}(X=0) = \binom{2}{0} (0.5)^0 (0.5)^2 = 1 \times 1 \times 0.25 = 0.25 = \frac{1}{4} \] \[ \mathbb{P}(X=1) = \binom{2}{1} (0.5)^1 (0.5)^1 = 2 \times 0.5 \times 0.5 = 0.5 = \frac{1}{2} \] \[ \mathbb{P}(X=2) = \binom{2}{2} (0.5)^2 (0.5)^0 = 1 \times 0.25 \times 1 = 0.25 = \frac{1}{4} \]

Now, we apply the law of total probability: \[ \mathbb{P}(U_1 + U_2 > X) = \sum_{k=0}^{2} \mathbb{P}(U_1 + U_2 > X | X=k) \mathbb{P}(X=k) \] \[ = \mathbb{P}(U_1 + U_2 > 0)\mathbb{P}(X=0) + \mathbb{P}(U_1 + U_2 > 1)\mathbb{P}(X=1) + \mathbb{P}(U_1 + U_2 > 2)\mathbb{P}(X=2) \]

Let's calculate each conditional probability. \( U_1 \) and \( U_2 \) are coordinates in a unit square in the \( u_1u_2 \)-plane. The total area is 1.

Case 1: \( \mathbb{P}(U_1 + U_2 > 0) \).

Since \( U_1, U_2 \in (0,1) \), their sum is always greater than 0. So, \( \mathbb{P}(U_1 + U_2 > 0) = 1 \).


Case 2: \( \mathbb{P}(U_1 + U_2 > 1) \).

This is 1 minus the probability \( \mathbb{P}(U_1 + U_2 \leq 1) \). The region \( u_1 + u_2 \leq 1 \) within the unit square is a triangle with vertices at (0,0), (1,0), and (0,1). The area of this triangle is \( \frac{1}{2} \times base \times height = \frac{1}{2} \times 1 \times 1 = \frac{1}{2} \).

So, \( \mathbb{P}(U_1 + U_2 \leq 1) = \frac{1}{2} \).

Therefore, \( \mathbb{P}(U_1 + U_2 > 1) = 1 - \frac{1}{2} = \frac{1}{2} \).


Case 3: \( \mathbb{P}(U_1 + U_2 > 2) \).

The maximum value of \( U_1 + U_2 \) is \( 1 + 1 = 2 \). The event \( U_1 + U_2 > 2 \) is impossible for continuous variables. The probability is 0.


Now substitute these probabilities back into the main equation: \[ \mathbb{P}(U_1 + U_2 > X) = (1) \left(\frac{1}{4}\right) + \left(\frac{1}{2}\right) \left(\frac{1}{2}\right) + (0) \left(\frac{1}{4}\right) \] \[ = \frac{1}{4} + \frac{1}{4} + 0 = \frac{2}{4} = \frac{1}{2} \]

Finally, we need to calculate \( 36 \mathbb{P}(U_1 + U_2 > X) \): \[ 36 \times \frac{1}{2} = 18 \]

Step 4: Final Answer:

The value of the expression is 18. Quick Tip: For problems involving sums of uniform random variables, visualizing the sample space as a unit square (for two variables) or a unit cube (for three) is extremely helpful. Probabilities correspond to areas or volumes within this space.


Question 50:

A drawer contains 5 pairs of shoes of different sizes. Assume that all 10 shoes are distinguishable. A person selects 5 shoes from the drawer at random. Then the probability that there are exactly 2 complete pairs of shoes among these 5 shoes is equal to ________________
(round off to 2 decimal places)

Correct Answer: 0.24 (Range 0.23 to 0.25)
View Solution




Step 1: Understanding the Concept:

This is a problem of combinatorial probability. We need to find the number of ways to select 5 shoes such that there are exactly two complete pairs, and divide it by the total number of ways to select 5 shoes from 10.


Step 2: Key Formula or Approach:

The probability of an event is given by the ratio of the number of favorable outcomes to the total number of possible outcomes.
\[ Probability = \frac{Number of Favorable Outcomes}{Total Number of Possible Outcomes} \]
We will use combinations, denoted as \( \binom{n}{k} = \frac{n!}{k!(n-k)!} \), to count the outcomes.


Step 3: Detailed Explanation:

Total Number of Possible Outcomes:

We are selecting 5 shoes from a total of 10 distinguishable shoes. The total number of ways to do this is: \[ Total ways = \binom{10}{5} = \frac{10!}{5!(10-5)!} = \frac{10 \times 9 \times 8 \times 7 \times 6}{5 \times 4 \times 3 \times 2 \times 1} = 2 \times 9 \times 2 \times 7 = 252 \]

Number of Favorable Outcomes:

We want to select exactly 2 complete pairs among the 5 shoes. This means we have 4 shoes forming 2 pairs, and 1 shoe that does not form a pair with any other selected shoe.


(A) Choose the 2 pairs: There are 5 pairs of shoes in total. The number of ways to choose 2 of these pairs is:
\[ \binom{5}{2} = \frac{5!}{2!3!} = \frac{5 \times 4}{2 \times 1} = 10 \]
This selection gives us 4 shoes.

(B) Choose the remaining single shoe: We need to select one more shoe. We have already selected 2 pairs, so 3 pairs (6 shoes) remain. The 5th shoe must be chosen from these remaining 6 shoes to avoid forming a third pair. The number of ways to choose 1 shoe from these 6 is:
\[ \binom{6}{1} = 6 \]

The total number of favorable outcomes is the product of the number of ways for each step: \[ Favorable ways = \binom{5}{2} \times \binom{6}{1} = 10 \times 6 = 60 \]

Calculate the Probability:
\[ Probability = \frac{Favorable ways}{Total ways} = \frac{60}{252} \]
To simplify the fraction, we can divide both the numerator and denominator by their greatest common divisor. \[ \frac{60}{252} = \frac{12 \times 5}{12 \times 21} = \frac{5}{21} \]

Convert to Decimal:

To round off to 2 decimal places, we perform the division: \[ \frac{5}{21} \approx 0.238095... \]
Rounding to 2 decimal places gives 0.24.


Step 4: Final Answer:

The probability is \( \frac{5}{21} \), which is approximately 0.24. Quick Tip: In combinatorics problems, break down the selection process into a sequence of distinct steps. For "exactly" type problems, ensure your final selection step doesn't accidentally satisfy a different condition (like choosing the 5th shoe to form a 3rd pair).


Question 51:

Let \( C = \{(x, y) \in \mathbb{R}^2 : x^2 + y^2 = 12\} \) be a circle in the plane. Let \( (a,b) \) be the point on C which minimizes the distance to the point (1,2). Then \( b-a \) is ________________
(round off to 2 decimal places)

Correct Answer: 1.55 (Range 1.54 to 1.56)
View Solution




Step 1: Understanding the Concept:

We are asked to find the point on a given circle that is closest to a given external point. The shortest distance from a point to a circle lies along the line connecting the center of the circle and the external point.


Step 2: Key Formula or Approach:

The circle \( C \) is given by \( x^2 + y^2 = 12 \).

This is a circle centered at the origin \( O(0,0) \) with radius \( r = \sqrt{12} = 2\sqrt{3} \).

The external point is \( P(1,2) \).

The point \( (a,b) \) on the circle closest to \( P \) will lie on the line segment connecting \( O \) and \( P \).


Step 3: Detailed Explanation:

First, find the equation of the line passing through the center \( O(0,0) \) and the point \( P(1,2) \).

The slope of the line is \( m = \frac{2-0}{1-0} = 2 \).

The equation of the line is \( y = mx \), which is \( y = 2x \).


The point \( (a,b) \) must lie on both the circle and this line. Therefore, it must satisfy both equations:

(A) \( a^2 + b^2 = 12 \) (Equation of the circle)
(B) \( b = 2a \) (Equation of the line)

Substitute the second equation into the first: \[ a^2 + (2a)^2 = 12 \] \[ a^2 + 4a^2 = 12 \] \[ 5a^2 = 12 \] \[ a^2 = \frac{12}{5} \] \[ a = \pm \sqrt{\frac{12}{5}} = \pm \frac{2\sqrt{3}}{\sqrt{5}} \]
Since the point \( P(1,2) \) is in the first quadrant, the closest point on the circle must also be in the first quadrant. Thus, we take the positive value for \( a \). \[ a = \frac{2\sqrt{3}}{\sqrt{5}} \]

Now, find the corresponding value of \( b \): \[ b = 2a = 2 \left( \frac{2\sqrt{3}}{\sqrt{5}} \right) = \frac{4\sqrt{3}}{\sqrt{5}} \]
So the point is \( (a,b) = \left( \frac{2\sqrt{3}}{\sqrt{5}}, \frac{4\sqrt{3}}{\sqrt{5}} \right) \).


The question asks for the value of \( b - a \). \[ b - a = \frac{4\sqrt{3}}{\sqrt{5}} - \frac{2\sqrt{3}}{\sqrt{5}} = \frac{2\sqrt{3}}{\sqrt{5}} \]

To round this to two decimal places, we calculate its numerical value: \[ \frac{2\sqrt{3}}{\sqrt{5}} = 2\sqrt{\frac{3}{5}} = 2\sqrt{0.6} \]
Using \( \sqrt{0.6} \approx 0.7746 \): \[ b - a \approx 2 \times 0.7746 = 1.5492 \]
Rounding to two decimal places, we get 1.55.


Step 4: Final Answer:

The value of \( b-a \) is approximately 1.55. Quick Tip: For distance minimization problems involving a circle, the geometric approach is often much faster than using calculus methods like Lagrange multipliers. The shortest (and longest) distance to a point outside a circle always lies on the line passing through the point and the circle's center.


Question 52:

Let \( f: \mathbb{R} \to \mathbb{R} \) and \( g: \mathbb{R} \to \mathbb{R} \) be given by \( f(x) = x^3 + 2x^2 - 15x \) and \( g(x) = x \) respectively. Let \( x_0 \) be the smallest strictly positive number such that \( f(x_0) = 0 \). Then the area of the region enclosed by the graphs of \( f \) and \( g \) between the lines \( x=0 \) and \( x = x_0 \) is ________________

Correct Answer: 33.75 (Range 33.70 to 33.80)
View Solution




Step 1: Understanding the Concept:

We need to find the area between two curves, \( y=f(x) \) and \( y=g(x) \), over a specified interval \( [0, x_0] \). First, we must determine the upper limit of integration, \( x_0 \), by finding the smallest positive root of \( f(x) = 0 \). Then, we will compute the definite integral of the absolute difference between the two functions.


Step 2: Key Formula or Approach:

The area \( A \) between two curves \( y=f(x) \) and \( y=g(x) \) from \( x=a \) to \( x=b \) is given by: \[ A = \int_{a}^{b} |f(x) - g(x)| \,dx \]

Step 3: Detailed Explanation:

Find the limit of integration \( x_0 \):

We are given that \( x_0 \) is the smallest strictly positive number such that \( f(x_0) = 0 \). \[ f(x) = x^3 + 2x^2 - 15x = 0 \]
Factor out \( x \): \[ x(x^2 + 2x - 15) = 0 \]
Factor the quadratic: \[ x(x+5)(x-3) = 0 \]
The roots are \( x = -5, 0, 3 \). The smallest strictly positive root is \( x_0 = 3 \).
So, we need to find the area between the curves from \( x=0 \) to \( x=3 \).


Set up the area integral:

The area \( A \) is given by \( \int_{0}^{3} |f(x) - g(x)| \,dx \).
Let \( h(x) = f(x) - g(x) \). \[ h(x) = (x^3 + 2x^2 - 15x) - x = x^3 + 2x^2 - 16x \]
To evaluate the integral with the absolute value, we must determine the sign of \( h(x) \) on the interval \( (0, 3) \). \[ h(x) = x(x^2 + 2x - 16) \]
For \( x \in (0,3) \), the term \( x \) is positive. We need to check the sign of \( x^2 + 2x - 16 \).
The roots of \( x^2 + 2x - 16 = 0 \) are given by the quadratic formula: \[ x = \frac{-2 \pm \sqrt{2^2 - 4(1)(-16)}}{2(1)} = \frac{-2 \pm \sqrt{4+64}}{2} = \frac{-2 \pm \sqrt{68}}{2} = -1 \pm \sqrt{17} \]
Since \( \sqrt{16}=4 \) and \( \sqrt{25}=5 \), \( \sqrt{17} \approx 4.12 \).
The positive root is \( -1 + \sqrt{17} \approx 3.12 \).
The quadratic \( x^2 + 2x - 16 \) is a parabola opening upwards, so it is negative between its roots. Since the interval \( (0, 3) \) is entirely within the roots \( (-1-\sqrt{17}, -1+\sqrt{17}) \), \( x^2 + 2x - 16 \) is negative on \( (0, 3) \).
Therefore, \( h(x) = x(negative number) \) is negative for \( x \in (0, 3) \).

This means \( |h(x)| = -h(x) \) on this interval.


Calculate the definite integral:
\[ A = \int_{0}^{3} -(x^3 + 2x^2 - 16x) \,dx = \int_{0}^{3} (-x^3 - 2x^2 + 16x) \,dx \] \[ A = \left[ -\frac{x^4}{4} - \frac{2x^3}{3} + \frac{16x^2}{2} \right]_{0}^{3} = \left[ -\frac{x^4}{4} - \frac{2x^3}{3} + 8x^2 \right]_{0}^{3} \] \[ A = \left( -\frac{3^4}{4} - \frac{2(3)^3}{3} + 8(3)^2 \right) - (0) \] \[ A = -\frac{81}{4} - \frac{2(27)}{3} + 8(9) \] \[ A = -20.25 - 18 + 72 \] \[ A = -38.25 + 72 = 33.75 \]

Step 4: Final Answer:

The area of the region is 33.75. Quick Tip: When calculating the area between curves, always check the sign of \( f(x) - g(x) \) over the integration interval. If it changes sign, you must split the integral into multiple parts. A quick test with a point inside the interval (e.g., x=1) can often determine the sign if there are no roots of \( f(x)-g(x)=0 \) within the interval.


Question 53:

Let V be the volume of the region \[ \{ (x,y,z) \in \mathbb{R}^3 : x^2 + y^2 + \frac{z^2}{4} \leq 1 and |z| \leq 1 \}. \]
Then \( \frac{V}{\pi} \) is equal to ________________

Correct Answer: 1.83 (Range 1.82 to 1.84)
View Solution




Step 1: Understanding the Concept:

We need to find the volume of a solid defined by two inequalities. The first inequality, \( x^2 + y^2 + \frac{z^2}{4} \leq 1 \), describes a solid ellipsoid centered at the origin. The second inequality, \( |z| \leq 1 \), which is equivalent to \( -1 \leq z \leq 1 \), restricts the solid to a specific range along the z-axis. We can compute this volume using the method of slicing (disk method).


Step 2: Key Formula or Approach:

The volume \( V \) of a solid can be found by integrating the cross-sectional area \( A(z) \) along the z-axis from \( z=a \) to \( z=b \). \[ V = \int_{a}^{b} A(z) \,dz \]

Step 3: Detailed Explanation:

The solid is defined by \( x^2 + y^2 + \frac{z^2}{4} \leq 1 \) and \( -1 \leq z \leq 1 \).

Let's find the cross-sectional area \( A(z) \) for a fixed value of \( z \) in the interval \( [-1, 1] \).
The inequality can be rearranged as: \[ x^2 + y^2 \leq 1 - \frac{z^2}{4} \]
For a fixed \( z \), this describes a disk in the xy-plane centered at the origin with radius \( r \) where \( r^2 = 1 - \frac{z^2}{4} \).
The area of this disk is \( A(z) = \pi r^2 \). \[ A(z) = \pi \left( 1 - \frac{z^2}{4} \right) \]

Now, we integrate this area function from \( z = -1 \) to \( z = 1 \) to find the volume \( V \). \[ V = \int_{-1}^{1} A(z) \,dz = \int_{-1}^{1} \pi \left( 1 - \frac{z^2}{4} \right) \,dz \]
Since the integrand \( \pi \left( 1 - \frac{z^2}{4} \right) \) is an even function of \( z \), we can simplify the integral: \[ V = 2 \int_{0}^{1} \pi \left( 1 - \frac{z^2}{4} \right) \,dz \] \[ V = 2\pi \int_{0}^{1} \left( 1 - \frac{z^2}{4} \right) \,dz \]
Now, compute the integral: \[ V = 2\pi \left[ z - \frac{z^3}{3 \times 4} \right]_{0}^{1} = 2\pi \left[ z - \frac{z^3}{12} \right]_{0}^{1} \] \[ V = 2\pi \left[ \left( 1 - \frac{1^3}{12} \right) - (0) \right] \] \[ V = 2\pi \left( 1 - \frac{1}{12} \right) = 2\pi \left( \frac{11}{12} \right) = \frac{22\pi}{12} = \frac{11\pi}{6} \]

The question asks for the value of \( \frac{V}{\pi} \). \[ \frac{V}{\pi} = \frac{11\pi/6}{\pi} = \frac{11}{6} \]

To get the numerical answer, we perform the division: \[ \frac{11}{6} = 1.8333... \]
Rounding this to two decimal places gives 1.83.


Step 4: Final Answer:

The value of \( \frac{V}{\pi} \) is approximately 1.83. Quick Tip: The method of slicing is a powerful tool for finding volumes of solids, especially those with rotational symmetry or easily described cross-sections like ellipsoids, cones, and pyramids. Always identify the shape of the cross-section first, find its area as a function of one variable, and then integrate that area function over the appropriate interval.


Question 54:

Suppose \( X_1, X_2, \dots, X_{10}, Y_1, Y_2, \dots, Y_{10} \) are independent random variables, where \( X_i \sim N(0, \sigma^2) \) and \( Y_i \sim N(0, 3\sigma^2) \) for \( i = 1, 2, \dots, 10 \). The observables are \( D_1, \dots, D_{10} \), where \( D_i \) denotes the Euclidean distance between the points \( (X_i, Y_i, 0) \) and \( (0, 0, 5) \) for \( i = 1, 2, \dots, 10 \). If the observed value of \( \sum_{i=1}^{10} D_i^2 \) is equal to 1050, then the method of moments estimate of \( \sigma^2 \) is equal to ________________
(answer in integer)

Correct Answer: 20
View Solution




Step 1: Understanding the Concept:

We need to find the method of moments (MOM) estimate for the parameter \( \sigma^2 \). The MOM involves equating the first sample moment to the first population moment and solving for the parameter. The observables are \( D_i \), but the given data is related to \( D_i^2 \), so we will use the moments of \( D_i^2 \).


Step 2: Key Formula or Approach:

1. Define the observable random variable and find its expected value (the population moment) in terms of the parameter \( \sigma^2 \).

2. Calculate the sample moment from the given data.

3. Equate the population moment to the sample moment to find the estimate.


Step 3: Detailed Explanation:

First, let's express \( D_i^2 \) in terms of \( X_i \) and \( Y_i \). The Euclidean distance \( D_i \) between \( (X_i, Y_i, 0) \) and \( (0, 0, 5) \) is: \[ D_i = \sqrt{(X_i-0)^2 + (Y_i-0)^2 + (0-5)^2} = \sqrt{X_i^2 + Y_i^2 + 25} \]
Squaring both sides gives: \[ D_i^2 = X_i^2 + Y_i^2 + 25 \]

Now, we find the first population moment, \( E[D_i^2] \). \[ E[D_i^2] = E[X_i^2 + Y_i^2 + 25] \]
By linearity of expectation: \[ E[D_i^2] = E[X_i^2] + E[Y_i^2] + E[25] = E[X_i^2] + E[Y_i^2] + 25 \]
We know that for a random variable \( Z \) with mean \( \mu \) and variance \( \tau^2 \), \( Var(Z) = E[Z^2] - (E[Z])^2 \).
For \( X_i \sim N(0, \sigma^2) \), we have \( E[X_i] = 0 \) and \( Var(X_i) = \sigma^2 \). \[ \sigma^2 = E[X_i^2] - (0)^2 \implies E[X_i^2] = \sigma^2 \]
For \( Y_i \sim N(0, 3\sigma^2) \), we have \( E[Y_i] = 0 \) and \( Var(Y_i) = 3\sigma^2 \). \[ 3\sigma^2 = E[Y_i^2] - (0)^2 \implies E[Y_i^2] = 3\sigma^2 \]
Substituting these back into the expression for \( E[D_i^2] \): \[ E[D_i^2] = \sigma^2 + 3\sigma^2 + 25 = 4\sigma^2 + 25 \]

Next, we calculate the first sample moment of \( D_i^2 \). The sample mean is: \[ \overline{D^2} = \frac{1}{10} \sum_{i=1}^{10} D_i^2 \]
Given that \( \sum_{i=1}^{10} D_i^2 = 1050 \): \[ \overline{D^2} = \frac{1050}{10} = 105 \]

For the method of moments, we equate the population moment to the sample moment: \[ E[D_i^2] = \overline{D^2} \] \[ 4\sigma^2 + 25 = 105 \]
Now, we solve for the estimate of \( \sigma^2 \), which we denote as \( \hat{\sigma}^2 \). \[ 4\hat{\sigma}^2 = 105 - 25 \] \[ 4\hat{\sigma}^2 = 80 \] \[ \hat{\sigma}^2 = \frac{80}{4} = 20 \]

Step 4: Final Answer:

The method of moments estimate of \( \sigma^2 \) is 20. Quick Tip: The method of moments is a straightforward estimation technique. Remember to identify the random variable whose sample data is provided (here, it's effectively \( D_i^2 \)) and match its sample mean with its theoretical mean (expected value).


Question 55:

Consider a sequence of independent Bernoulli trials with success probability \( p = \frac{1}{7} \). Then the expected number of trials required to get two consecutive successes for the first time is equal to ________________

Correct Answer: 56
View Solution




Step 1: Understanding the Concept:

This is a classic problem on expected values in stochastic processes. We can solve it by setting up a system of linear equations based on conditional expectations from different states of the process. The states are defined by the outcome of the most recent trial(s).


Step 2: Key Formula or Approach:

Let \( E \) be the expected number of trials required to get two consecutive successes (SS). We can define states based on the progress towards this goal:

State 0: We are at the start, or the last trial was a failure (F). Let \( E_0 \) be the expected number of additional trials needed from here. This is our target value, \( E = E_0 \).
State 1: The last trial was a success (S). Let \( E_1 \) be the expected number of additional trials needed from here.

We will use the law of total expectation to create equations for \( E_0 \) and \( E_1 \).


Step 3: Detailed Explanation:

Let S denote success and F denote failure. The probability of success is \( p = 1/7 \), and the probability of failure is \( q = 1 - p = 6/7 \).


From State 0:
We perform one trial.

If we get a failure (with probability \( q \)), we have spent 1 trial and are back in State 0.
If we get a success (with probability \( p \)), we have spent 1 trial and move to State 1.

So, the equation for \( E_0 \) is: \[ E_0 = 1 + q \cdot E_0 + p \cdot E_1 \]

From State 1:
We perform one trial.

If we get a failure (with probability \( q \)), we have spent 1 trial, the streak is broken, and we return to State 0.
If we get a success (with probability \( p \)), we have spent 1 trial and have achieved our goal (SS). The process ends, requiring 0 more trials.

So, the equation for \( E_1 \) is: \[ E_1 = 1 + q \cdot E_0 + p \cdot 0 = 1 + q E_0 \]

Now we have a system of two equations:
1. \( E_0 = 1 + q E_0 + p E_1 \)
2. \( E_1 = 1 + q E_0 \)

Substitute equation (2) into (1): \[ E_0 = 1 + q E_0 + p(1 + q E_0) \] \[ E_0 = 1 + q E_0 + p + pq E_0 \]
Rearrange to solve for \( E_0 \): \[ E_0 - q E_0 - pq E_0 = 1 + p \] \[ E_0(1 - q - pq) = 1 + p \]
Since \( 1 - q = p \): \[ E_0(p - pq) = 1 + p \] \[ E_0 p(1-q) = 1 + p \]
Since \( 1 - q = p \): \[ E_0 p^2 = 1 + p \] \[ E_0 = \frac{1+p}{p^2} \]

Now, substitute the value \( p = 1/7 \): \[ E = E_0 = \frac{1 + 1/7}{(1/7)^2} = \frac{8/7}{1/49} = \frac{8}{7} \times 49 = 8 \times 7 = 56 \]

Step 4: Final Answer:

The expected number of trials is 56. Quick Tip: For problems asking for the expected time to reach a certain pattern in a sequence of trials, the state-based conditional expectation method is very effective. Clearly define the states and write down the transition equations by considering the outcome of the next trial.


Question 56:

Let X be a real valued random variable with \( E(X)=1 \), \( E(X^2)=4 \), \( E(X^4)=16 \). Then \( E(X^3) \) is equal to ________________

Correct Answer: 4
View Solution




Step 1: Understanding the Concept:

The problem provides the first, second, and fourth moments of a random variable \( X \) and asks for the third moment. The relationship between the given moments \( E(X^2) \) and \( E(X^4) \) is the key to solving this problem.


Step 2: Key Formula or Approach:

We will analyze the variance of the random variable \( Y = X^2 \). The variance of any random variable \( Y \) is given by \( Var(Y) = E[Y^2] - (E[Y])^2 \). A key property is that if \( Var(Y) = 0 \), then \( Y \) must be a constant almost surely, equal to its expected value.


Step 3: Detailed Explanation:

Let's define a new random variable \( Y = X^2 \). We are given information about the moments of \( X \), which we can use to find the moments of \( Y \).
The expected value of \( Y \) is: \[ E[Y] = E[X^2] = 4 \]
The expected value of \( Y^2 \) is: \[ E[Y^2] = E[(X^2)^2] = E[X^4] = 16 \]
Now, let's calculate the variance of \( Y \): \[ Var(Y) = E[Y^2] - (E[Y])^2 = 16 - (4)^2 = 16 - 16 = 0 \]
Since the variance of \( Y \) is 0, \( Y \) must be a constant almost surely. This constant is equal to its expected value. \[ Y = E[Y] \implies X^2 = 4 \]
This means that the random variable \( X \) can only take two possible values: 2 and -2.


Let \( \mathbb{P}(X=2) = p \). Then, since \( X \) can only be 2 or -2, \( \mathbb{P}(X=-2) = 1-p \).

We can use the given first moment, \( E[X]=1 \), to find the value of \( p \). \[ E[X] = (2) \cdot \mathbb{P}(X=2) + (-2) \cdot \mathbb{P}(X=-2) = 1 \] \[ 2p - 2(1-p) = 1 \] \[ 2p - 2 + 2p = 1 \] \[ 4p = 3 \implies p = \frac{3}{4} \]
So, the probability distribution of \( X \) is: \[ \mathbb{P}(X=2) = \frac{3}{4} \quad and \quad \mathbb{P}(X=-2) = 1 - \frac{3}{4} = \frac{1}{4} \]
We can verify that this distribution gives the correct second and fourth moments: \( E[X^2] = (2)^2(\frac{3}{4}) + (-2)^2(\frac{1}{4}) = 4(\frac{3}{4}) + 4(\frac{1}{4}) = 3 + 1 = 4 \). (Correct) \( E[X^4] = (2)^4(\frac{3}{4}) + (-2)^4(\frac{1}{4}) = 16(\frac{3}{4}) + 16(\frac{1}{4}) = 12 + 4 = 16 \). (Correct)

Finally, we can calculate the third moment, \( E[X^3] \): \[ E[X^3] = (2)^3 \cdot \mathbb{P}(X=2) + (-2)^3 \cdot \mathbb{P}(X=-2) \] \[ E[X^3] = (8) \left(\frac{3}{4}\right) + (-8) \left(\frac{1}{4}\right) \] \[ E[X^3] = 6 - 2 = 4 \]

Step 4: Final Answer:

The value of \( E(X^3) \) is 4. Quick Tip: When given several moments of a random variable, check for simple relationships between them. A condition like \( E[Y^2] = (E[Y])^2 \) is a strong indicator that \( Y \) has zero variance and is therefore a constant. This can greatly simplify the problem by revealing the structure of the underlying random variable.


Question 57:

Let \( X_1, X_2, X_3, X_4 \) be a random sample from a continuous distribution with probability density function

where \( \theta > 0 \) is an unknown parameter. It is known that \( X_{(1)} = \min\{X_1, X_2, X_3, X_4\} \) is a complete sufficient statistic for \( \theta \). If the observed values are \( x_1 = 15, x_2 = 11, x_3 = 10, x_4 = 17 \), the uniformly minimum variance unbiased estimate of \( \theta^2 \) is equal to ________________

Correct Answer: 75
View Solution




Step 1: Understanding the Concept:

We are asked to find the Uniformly Minimum Variance Unbiased Estimator (UMVUE) of \( \theta^2 \). According to the Lehmann-Scheffé theorem, if a statistic \( S \) is complete and sufficient for a parameter \( \theta \), then any function \( g(S) \) that is an unbiased estimator for a quantity \( \tau(\theta) \) is the UMVUE of \( \tau(\theta) \). Here, \( S = X_{(1)} \) and \( \tau(\theta) = \theta^2 \).


Step 2: Key Formula or Approach:

1. Find the probability distribution (PDF) of the complete sufficient statistic \( X_{(1)} \).
2. Find a function \( g(X_{(1)}) \) such that its expected value is equal to \( \theta^2 \). That is, \( E[g(X_{(1)})] = \theta^2 \).
3. This function \( g(X_{(1)}) \) will be the UMVUE for \( \theta^2 \).
4. Substitute the observed value of \( X_{(1)} \) into the UMVUE to get the estimate.


Step 3: Detailed Explanation:

1. Distribution of \( X_{(1)} \):
First, find the CDF of \( X \), \( F(x) \), for \( x > \theta \): \[ F(x) = \int_{\theta}^{x} 2\theta^2 t^{-3} dt = 2\theta^2 \left[ \frac{t^{-2}}{-2} \right]_{\theta}^{x} = 2\theta^2 \left( -\frac{1}{2x^2} - (-\frac{1}{2\theta^2}) \right) = 1 - \frac{\theta^2}{x^2} \]
The CDF of \( X_{(1)} \) for a sample of size \( n=4 \) is \( F_{X_{(1)}}(y) = 1 - (1 - F(y))^n \). \[ F_{X_{(1)}}(y) = 1 - \left(1 - \left(1 - \frac{\theta^2}{y^2}\right)\right)^4 = 1 - \left(\frac{\theta^2}{y^2}\right)^4 = 1 - \frac{\theta^8}{y^8} \quad for y > \theta \]
The PDF of \( X_{(1)} \) is the derivative of its CDF: \[ f_{X_{(1)}}(y) = \frac{d}{dy} \left( 1 - \theta^8 y^{-8} \right) = -(-8) \theta^8 y^{-9} = 8\theta^8 y^{-9} \quad for y > \theta \]

2. Find an unbiased estimator of \( \theta^2 \):
Let's find the expectation of a power of \( X_{(1)} \), say \( E[X_{(1)}^k] \). \[ E[X_{(1)}^k] = \int_{\theta}^{\infty} y^k f_{X_{(1)}}(y) dy = \int_{\theta}^{\infty} y^k (8\theta^8 y^{-9}) dy = 8\theta^8 \int_{\theta}^{\infty} y^{k-9} dy \]
This integral converges for \( k-9 < -1 \), i.e., \( k < 8 \). \[ E[X_{(1)}^k] = 8\theta^8 \left[ \frac{y^{k-8}}{k-8} \right]_{\theta}^{\infty} = 8\theta^8 \left( 0 - \frac{\theta^{k-8}}{k-8} \right) = \frac{8\theta^k}{8-k} \]
We are looking for an estimator for \( \theta^2 \). Let's try to find a constant \( c \) such that \( E[c X_{(1)}^2] = \theta^2 \). We use \( k=2 \). \[ E[X_{(1)}^2] = \frac{8\theta^2}{8-2} = \frac{8\theta^2}{6} = \frac{4}{3}\theta^2 \]
So, \( E[c X_{(1)}^2] = c E[X_{(1)}^2] = c \frac{4}{3}\theta^2 \).
To make this estimator unbiased for \( \theta^2 \), we set \( c \frac{4}{3} = 1 \), which gives \( c = \frac{3}{4} \).
Thus, the UMVUE for \( \theta^2 \) is \( g(X_{(1)}) = \frac{3}{4}X_{(1)}^2 \).

3. Calculate the estimate:
The observed values are \( x_1=15, x_2=11, x_3=10, x_4=17 \).
The observed value of the statistic \( X_{(1)} \) is: \[ x_{(1)} = \min\{15, 11, 10, 17\} = 10 \]
Now we substitute this value into our UMVUE formula: \[ Estimate of \theta^2 = \frac{3}{4} x_{(1)}^2 = \frac{3}{4} (10)^2 = \frac{3}{4} \times 100 = 75 \]

Step 4: Final Answer:

The uniformly minimum variance unbiased estimate of \( \theta^2 \) is 75. Quick Tip: The Lehmann-Scheffé theorem is a powerful tool for finding UMVUEs. The main steps are identifying a complete sufficient statistic and then finding a function of that statistic which is unbiased for the parameter of interest. For scale parameter families like this one, moments of the sufficient statistic are often a good starting point.


Question 58:

Let X be a random variable with probability density function

Let \( \delta \) denote the conditional expectation of \( X \) given that \( X \leq \frac{1}{2} \).
Then the value of \( 80\delta \) is equal to ________________

Correct Answer: 30
View Solution




Step 1: Understanding the Concept:

We need to calculate the conditional expectation of a continuous random variable \( X \) given an event \( A \). The event here is \( A = \{X \leq 1/2\} \).


Step 2: Key Formula or Approach:

The conditional expectation of \( X \) given an event \( A \) is defined as: \[ E[X | A] = \frac{\int_A x f(x) \,dx}{\mathbb{P}(A)} \]
where \( \mathbb{P}(A) = \int_A f(x) \,dx \).
In this case, \( A \) corresponds to the interval \( (0, 1/2] \) since the support of \( f(x) \) is \( (0, 1) \).
So, \( \delta = E[X | X \leq 1/2] = \frac{\int_0^{1/2} x f(x) \,dx}{\int_0^{1/2} f(x) \,dx} \).


Step 3: Detailed Explanation:

First, we calculate the denominator, which is the probability of the event \( \mathbb{P}(X \leq 1/2) \). \[ \mathbb{P}(X \leq 1/2) = \int_0^{1/2} f(x) \,dx = \int_0^{1/2} 3x^2 \,dx \] \[ = \left[ x^3 \right]_0^{1/2} = \left(\frac{1}{2}\right)^3 - 0^3 = \frac{1}{8} \]

Next, we calculate the numerator. \[ \int_0^{1/2} x f(x) \,dx = \int_0^{1/2} x (3x^2) \,dx = \int_0^{1/2} 3x^3 \,dx \] \[ = 3 \left[ \frac{x^4}{4} \right]_0^{1/2} = 3 \left( \frac{(1/2)^4}{4} - 0 \right) = 3 \left( \frac{1/16}{4} \right) = \frac{3}{64} \]

Now, we can compute \( \delta \). \[ \delta = \frac{numerator}{denominator} = \frac{3/64}{1/8} = \frac{3}{64} \times \frac{8}{1} = \frac{24}{64} = \frac{3}{8} \]

Finally, we calculate the required value of \( 80\delta \). \[ 80\delta = 80 \times \frac{3}{8} = 10 \times 3 = 30 \]

Step 4: Final Answer:

The value of \( 80\delta \) is 30. Quick Tip: For conditional expectation problems with continuous variables, remember the definition involves two integrals. The denominator is the probability of the conditioning event, which acts as a normalizing constant for the new "truncated" distribution.


Question 59:

Let \( X_1, X_2, \dots, X_7 \) be i.i.d. continuous random variables with median \( \theta \). If \( X_{(1)} < X_{(2)} < \dots < X_{(7)} \) are the corresponding order statistics, then \( \mathbb{P}(X_{(2)} > \theta) \) is equal to ________________
(round off to 3 decimal places).

Correct Answer: 0.063
View Solution




Step 1: Understanding the Concept:

The problem involves order statistics and the definition of a median. The median \( \theta \) of a continuous distribution is the value for which \( \mathbb{P}(X_i \leq \theta) = 0.5 \). We can rephrase the event \( X_{(2)} > \theta \) in terms of the number of observations that fall above or below the median.


Step 2: Key Formula or Approach:

Let's classify each random variable \( X_i \) as a 'success' if \( X_i > \theta \) and a 'failure' if \( X_i \leq \theta \). By definition of the median, the probability of success is \( p = \mathbb{P}(X_i > \theta) = 0.5 \).
The event \( X_{(2)} > \theta \) means that the second smallest observation is greater than the median. This can only happen if at most one observation is less than or equal to the median. In other words, out of the 7 observations, either 0 or 1 are \( \leq \theta \). This is equivalent to saying that either 7 or 6 observations are \( > \theta \).
Let \( S \) be the number of successes (i.e., the number of \( X_i \) such that \( X_i > \theta \)). \( S \) follows a binomial distribution \( Bin(n=7, p=0.5) \). We need to calculate \( \mathbb{P}(S=6) + \mathbb{P}(S=7) \).


Step 3: Detailed Explanation:

The number of observations greater than the median, \( S \), follows a Binomial distribution with parameters \( n=7 \) and \( p=0.5 \). The probability mass function (PMF) is: \[ \mathbb{P}(S=k) = \binom{n}{k} p^k (1-p)^{n-k} = \binom{7}{k} (0.5)^k (0.5)^{7-k} = \binom{7}{k} (0.5)^7 \]
The event \( X_{(2)} > \theta \) is equivalent to the event that there are 6 or 7 observations greater than \( \theta \), i.e., \( S \geq 6 \). \[ \mathbb{P}(X_{(2)} > \theta) = \mathbb{P}(S \geq 6) = \mathbb{P}(S=6) + \mathbb{P}(S=7) \]
Calculate each term: \[ \mathbb{P}(S=6) = \binom{7}{6} (0.5)^7 = 7 \times \frac{1}{128} = \frac{7}{128} \] \[ \mathbb{P}(S=7) = \binom{7}{7} (0.5)^7 = 1 \times \frac{1}{128} = \frac{1}{128} \]
Sum the probabilities: \[ \mathbb{P}(X_{(2)} > \theta) = \frac{7}{128} + \frac{1}{128} = \frac{8}{128} = \frac{1}{16} \]

To provide the answer in the required format, we convert the fraction to a decimal: \[ \frac{1}{16} = 0.0625 \]
Rounding to 3 decimal places gives 0.063.


Step 4: Final Answer:

The probability is \( \frac{1}{16} = 0.0625 \), which rounds to 0.063. Quick Tip: Problems involving order statistics relative to a quantile (like the median) can often be simplified by converting them into a binomial probability problem. Classify each observation as being above or below the quantile, and then count the number of ways the desired arrangement of order statistics can occur.


Question 60:

Suppose \( (X,Y) \) has the \( N_2(3, 0, 4, 1, 0.5) \) distribution. Then \( 4 Cov(X + Y, Y^3) \) is equal to ________________

Correct Answer: 24
View Solution




Step 1: Understanding the Concept:

We need to compute the covariance involving a sum of random variables and a power of one of them, where the variables follow a bivariate normal distribution. We will use properties of covariance and moments of normal distributions.


Step 2: Key Formula or Approach:

1. Use the bilinearity of covariance: \( Cov(A+B, C) = Cov(A,C) + Cov(B,C) \).

2. Use the definition of covariance: \( Cov(U,V) = E[UV] - E[U]E[V] \).

3. Use Isserlis' theorem (or Wick's theorem) for moments of zero-mean multivariate normal variables. For a zero-mean normal vector \( (W_1, W_2, W_3, W_4) \), \( E[W_1 W_2 W_3 W_4] = E[W_1W_2]E[W_3W_4] + E[W_1W_3]E[W_2W_4] + E[W_1W_4]E[W_2W_3] \).


Step 3: Detailed Explanation:

The parameters of the bivariate normal distribution \( N_2(\mu_X, \mu_Y, \sigma_X^2, \sigma_Y^2, \rho) \) are: \( \mu_X = E[X] = 3 \), \( \mu_Y = E[Y] = 0 \), \( \sigma_X^2 = 4 \), \( \sigma_Y^2 = 1 \), \( \rho = 0.5 \).


First, expand the covariance term: \[ Cov(X+Y, Y^3) = Cov(X, Y^3) + Cov(Y, Y^3) \]

Calculate \( Cov(Y, Y^3) \):
Since \( Y \sim N(0, 1) \), it is a standard normal variable. \[ Cov(Y, Y^3) = E[Y \cdot Y^3] - E[Y]E[Y^3] = E[Y^4] - E[Y]E[Y^3] \]
For a standard normal distribution, all odd moments are zero. So, \( E[Y] = 0 \) and \( E[Y^3] = 0 \).
The even moments are given by \( E[Y^{2k}] = (2k-1)!! \).
For \( k=2 \), \( E[Y^4] = (2(2)-1)!! = 3!! = 3 \times 1 = 3 \). \[ Cov(Y, Y^3) = 3 - (0)(0) = 3 \]

Calculate \( Cov(X, Y^3) \): \[ Cov(X, Y^3) = E[X Y^3] - E[X]E[Y^3] \]
We know \( E[X]=3 \) and \( E[Y^3]=0 \), so \( Cov(X, Y^3) = E[X Y^3] \).
To calculate \( E[X Y^3] \), we work with the centered variables \( X_0 = X - \mu_X = X - 3 \) and \( Y_0 = Y - \mu_Y = Y - 0 = Y \). \[ E[X Y^3] = E[(X_0 + 3) Y^3] = E[X_0 Y^3] + 3E[Y^3] = E[X_0 Y^3] + 0 = E[X_0 Y^3] \]
Now we apply Isserlis' theorem to the zero-mean normal variables \( (X_0, Y, Y, Y) \). \[ E[X_0 Y Y Y] = E[X_0 Y]E[Y Y] + E[X_0 Y]E[Y Y] + E[X_0 Y]E[Y Y] \]
Note: This is because the three Y variables are identical, leading to three identical terms in the sum over pairings. \[ E[X_0 Y^3] = 3 E[X_0 Y] E[Y^2] \]
We need \( E[X_0 Y] \) and \( E[Y^2] \). \[ E[X_0 Y] = E[(X-\mu_X)(Y-\mu_Y)] = Cov(X,Y) \] \[ Cov(X,Y) = \rho \sigma_X \sigma_Y = 0.5 \times \sqrt{4} \times \sqrt{1} = 0.5 \times 2 \times 1 = 1 \] \[ E[Y^2] = Var(Y) + (E[Y])^2 = 1 + 0^2 = 1 \]
Substituting these values back: \[ E[X_0 Y^3] = 3 \times (1) \times (1) = 3 \]
So, \( Cov(X, Y^3) = 3 \).

Combine the results: \[ Cov(X+Y, Y^3) = Cov(X, Y^3) + Cov(Y, Y^3) = 3 + 3 = 6 \]

Finally, calculate the required expression: \[ 4 Cov(X + Y, Y^3) = 4 \times 6 = 24 \]

Step 4: Final Answer:

The value of \( 4 Cov(X + Y, Y^3) \) is 24. Quick Tip: For calculating moments and covariances of functions of bivariate normal variables, Isserlis' theorem (or Wick's theorem) is extremely useful. It allows breaking down expectations of products of normal variables into sums of products of pairwise expectations (covariances). Always remember to center the variables (subtract their means) before applying the theorem.

IIT JAM Questions

  • 1.
    A light beam given by \(\mathbf{E}(z, t) = E_{01} \sin(kz - \omega t)\hat{i} + E_{02} \sin(kz - \omega t + \frac{\pi}{6})\hat{j}\) passes through an ideal linear polarizer whose transmission axis is tilted by 60\(^\circ\) from x-axis (in x-y plane). If \(E_{01} = 4\) V/m and \(E_{02} = 2\) V/m, the electric field amplitude of the emerging light beam from the polarizer is ______ V/m. (up to two decimal places)


      • 2.
        If \(\left(\frac{1-i}{1+i}\right)^{n/2} = -1\), where \(i = \sqrt{-1}\), one possible value of n is

          • 2
          • 4
          • 6
          • 8

        • 3.
          Consider a chamber at room temperature (27 \(^\circ\)C) filled with a gas having a molecular diameter of 0.35 nm. The pressure (in Pascal) to which the chamber needs to be evacuated so that the molecules have a mean free path of 1 km is \rule{1cm{0.15mm} \(\times 10^{-5}\) Pa. (up to two decimal places)
          (Boltzmann constant \(k_B = 1.38 \times 10^{-23}\) J/K)}


            • 4.
              One kg of water at 27\(^\circ\)C is brought in contact with a heat reservoir kept at 37\(^\circ\)C. Upon reaching thermal equilibrium, this mass of water is brought in contact with another heat reservoir kept at 47\(^\circ\)C. The final temperature of water is 47\(^\circ\)C. The change in entropy of the whole system in this entire process is \rule{1cm{0.15mm} cal/K. (up to two decimal places)
              Take specific heat at constant pressure of water as 1 cal/(g K)}


                • 5.
                  In a two-level atomic system, the excited state is 0.2 eV above the ground state. Considering the Maxwell-Boltzmann distribution, the temperature at which 2% of the atoms will be in the excited state is _____ K. (up to two decimal places)
                  (Boltzmann constant \(k_B = 8.62 \times 10^{-5}\) eV/K)


                    • 6.
                      The ratio of the density of atoms between the (111) and (110) planes in a simple cubic (sc) lattice is ____. (up to two decimal places)

                        Fees Structure

                        Structure based on different categories

                        CategoriesState
                        General1500
                        sc750

                        In case of any inaccuracy, Notify Us! 

                        Comments


                        No Comments To Show