Maximum Likelihood Estimation For Uniform Distribution A Detailed Guide

by ADMIN 72 views

In statistical inference, estimating parameters of a distribution is a fundamental problem. When dealing with a uniform distribution, where all values within a given range are equally likely, the task becomes particularly interesting. Let's consider a scenario where we have a random sample x1,x2,...,xnx_1, x_2, ..., x_n drawn from a uniform distribution defined by the probability density function (pdf):

f(x)=1θ,0<x<θ,θ>0f(x) = \frac{1}{\theta}, \quad 0 < x < \theta, \quad \theta > 0

Here, θ\theta is the unknown parameter we aim to estimate. The challenge lies in finding a suitable estimator for θ\theta based on the observed sample. This article delves into the process of finding an estimator for θ\theta, with a focus on Maximum Likelihood Estimation (MLE). MLE is a widely used method for estimating parameters of a statistical model. It involves finding the parameter value that maximizes the likelihood function, which represents the probability of observing the given sample data.

Maximum Likelihood Estimation (MLE) Theory

To understand the approach for estimating θ\theta, it's crucial to grasp the core concept of Maximum Likelihood Estimation (MLE). MLE is a statistical method used to estimate the parameters of a probability distribution by maximizing the likelihood function. In simpler terms, MLE seeks to find the values of the parameters that make the observed data most probable. Let's delve deeper into the theoretical framework behind MLE, exploring its key principles and steps.

The Likelihood Function

The foundation of MLE is the likelihood function, which quantifies the plausibility of different parameter values given the observed data. For a random sample x1,x2,...,xnx_1, x_2, ..., x_n drawn from a distribution with probability density function (pdf) f(x;θ)f(x; \theta), where θ\theta is the parameter (or vector of parameters) to be estimated, the likelihood function is defined as:

L(θ;x1,x2,...,xn)=i=1nf(xi;θ)L(\theta; x_1, x_2, ..., x_n) = \prod_{i=1}^{n} f(x_i; \theta)

This equation represents the product of the pdf evaluated at each data point xix_i, with θ\theta as the parameter. The likelihood function essentially calculates the probability of observing the given data for different values of θ\theta. In our uniform distribution case, the likelihood function will be constructed based on the pdf f(x)=1θf(x) = \frac{1}{\theta}. We will see how the constraints of the uniform distribution (i.e., 0<x<θ0 < x < \theta) play a crucial role in defining the likelihood function.

Maximizing the Likelihood

The primary goal of MLE is to find the value of θ\theta that maximizes the likelihood function. This value, denoted as θ^\hat{\theta}, is the Maximum Likelihood Estimator (MLE) of θ\theta. In many cases, it is more convenient to work with the log-likelihood function, which is the natural logarithm of the likelihood function:

(θ;x1,x2,...,xn)=lnL(θ;x1,x2,...,xn)=i=1nlnf(xi;θ) \ell(\theta; x_1, x_2, ..., x_n) = \ln L(\theta; x_1, x_2, ..., x_n) = \sum_{i=1}^{n} \ln f(x_i; \theta)

Maximizing the log-likelihood function is equivalent to maximizing the likelihood function because the logarithm is a monotonically increasing function. Taking the logarithm simplifies the calculations, especially when dealing with products, as it transforms products into sums. The log-likelihood function for our uniform distribution will be derived and analyzed in the subsequent sections.

Steps in MLE

The process of finding the MLE typically involves the following steps:

  1. Formulate the likelihood function: This involves identifying the pdf of the distribution and constructing the likelihood function based on the sample data.
  2. Formulate the log-likelihood function: Taking the natural logarithm of the likelihood function simplifies the calculations.
  3. Differentiate the log-likelihood function: Calculate the derivative (or partial derivatives for multiple parameters) of the log-likelihood function with respect to the parameter(s).
  4. Set the derivative(s) to zero: Solve the equation(s) obtained by setting the derivative(s) to zero. The solutions are the critical points of the log-likelihood function.
  5. Verify the maximum: Check whether the critical point corresponds to a maximum (e.g., using the second derivative test). However, for certain distributions like the uniform distribution, the maximum might occur at the boundary of the parameter space, requiring a different approach.
  6. Determine the MLE: The value of the parameter at the maximum point is the MLE.\

The specific steps and techniques for maximizing the likelihood function can vary depending on the distribution and the complexity of the likelihood function. For some distributions, analytical solutions can be obtained by solving the equations derived from differentiation. However, for other distributions, numerical optimization methods may be required to find the MLE. In the case of the uniform distribution, we will see that the MLE is not found through standard differentiation but rather through careful consideration of the likelihood function's properties and the parameter space constraints.

Constructing the Likelihood Function for the Uniform Distribution

To find the MLE for θ\theta in the uniform distribution, the first step is to construct the likelihood function. Given the pdf:

f(x)=1θ,0<x<θf(x) = \frac{1}{\theta}, \quad 0 < x < \theta

and a random sample x1,x2,...,xnx_1, x_2, ..., x_n, the likelihood function is the product of the pdf evaluated at each sample point:

L(θ;x1,x2,...,xn)=i=1nf(xi;θ)=i=1n1θL(\theta; x_1, x_2, ..., x_n) = \prod_{i=1}^{n} f(x_i; \theta) = \prod_{i=1}^{n} \frac{1}{\theta}

This simplifies to:

L(θ)=1θnL(\theta) = \frac{1}{\theta^n}

However, this is not the complete picture. We must consider the support of the uniform distribution, which is 0<x<θ0 < x < \theta. This constraint implies that each xix_i must be less than θ\theta. Therefore, θ\theta must be greater than the maximum value in the sample. Let's denote the maximum value in the sample as:

x(n)=max(x1,x2,...,xn)x_{(n)} = \max(x_1, x_2, ..., x_n)

Now, we can express the likelihood function more accurately as:

L(θ)={1θnif θ>x(n)0if θx(n)L(\theta) = \begin{cases} \frac{1}{\theta^n} & \text{if } \theta > x_{(n)} \\ 0 & \text{if } \theta \leq x_{(n)} \end{cases}

This likelihood function is a crucial expression for finding the MLE of θ\theta. It highlights the interplay between the parameter θ\theta and the observed data. The function states that the likelihood is 1θn\frac{1}{\theta^n} if θ\theta is greater than the largest observation in the sample, and it is zero otherwise. The implication of this function's form on finding the MLE will be discussed in the next section. Understanding this construction is pivotal for determining how we can accurately estimate θ\theta from the sample data. The next step is to analyze this likelihood function to find the value of θ\theta that maximizes it.

Maximizing the Likelihood Function

Now that we have the likelihood function:

L(θ)={1θnif θ>x(n)0if θx(n)L(\theta) = \begin{cases} \frac{1}{\theta^n} & \text{if } \theta > x_{(n)} \\ 0 & \text{if } \theta \leq x_{(n)} \end{cases}

where x(n)=max(x1,x2,...,xn)x_{(n)} = \max(x_1, x_2, ..., x_n), we aim to find the value of θ\theta that maximizes L(θ)L(\theta). Unlike typical MLE problems where we take derivatives, in this case, we need to analyze the behavior of the likelihood function directly.

Analyzing the Likelihood Function

The likelihood function 1θn\frac{1}{\theta^n} is a decreasing function of θ\theta for θ>x(n)\theta > x_{(n)}. This means that as θ\theta increases, the likelihood decreases. However, we also know that L(θ)=0L(\theta) = 0 for θx(n)\theta \leq x_{(n)}. So, we want to find the smallest possible value of θ\theta that still allows L(θ)L(\theta) to be non-zero. In other words, we want to minimize θ\theta while ensuring that θ>x(n)\theta > x_{(n)}.

Finding the MLE

The smallest possible value for θ\theta that satisfies the condition θ>x(n)\theta > x_{(n)} is a value infinitesimally larger than x(n)x_{(n)}. However, since we are looking for the value that maximizes the likelihood, we consider the infimum of all such values. Therefore, the MLE of θ\theta, denoted as θ^\hat{\theta}, is:

θ^=x(n)=max(x1,x2,...,xn)\hat{\theta} = x_{(n)} = \max(x_1, x_2, ..., x_n)

This result is quite intuitive: the best estimate for the upper bound of the uniform distribution is the maximum value observed in the sample. If we chose any value smaller than x(n)x_{(n)}, the likelihood would be zero because we observed a data point larger than the supposed upper bound. This approach to finding the MLE highlights an important aspect of statistical estimation: sometimes, the solution requires a careful examination of the function's properties rather than a mechanical application of calculus.

Intuition Behind the Result

The result that the MLE of θ\theta is the maximum of the sample, x(n)x_{(n)}, makes intuitive sense. Consider the nature of the uniform distribution. All values between 0 and θ\theta are equally likely. If we observe a sample of values, the largest value in our sample provides a natural lower bound for our estimate of θ\theta. We can't have a θ\theta smaller than the largest observed value because that would make the observation impossible. Therefore, our estimate for θ\theta should be at least as large as the maximum observed value.

Since the likelihood function is decreasing in θ\theta beyond this point, we choose the smallest possible value for θ\theta that is still consistent with our observations. This leads us to the maximum order statistic, x(n)x_{(n)}, as the MLE. The fact that the MLE is the sample maximum also illustrates a crucial point about MLE: it's not always about finding a peak in a smooth curve. Sometimes, it's about finding the boundary of a feasible region, as is the case here. The next sections will delve into the properties of this estimator, such as its bias and how it performs as the sample size increases. This understanding is vital for assessing the reliability and accuracy of our estimate.

Properties of the Estimator

Now that we have found the MLE for θ\theta in the uniform distribution, which is the sample maximum x(n)x_{(n)}, it is essential to evaluate its properties. Understanding the properties of an estimator allows us to assess its reliability and performance. Key properties to consider include bias, variance, and consistency. Let's analyze these properties for our estimator θ^=x(n)\hat{\theta} = x_{(n)}.

Bias of the Estimator

The bias of an estimator is the difference between its expected value and the true value of the parameter being estimated. An estimator is said to be unbiased if its expected value is equal to the true parameter value. The bias, denoted as Bias(θ^)Bias(\hat{\theta}), is calculated as:

Bias(θ^)=E[θ^]θBias(\hat{\theta}) = E[\hat{\theta}] - \theta

To determine the bias of our estimator θ^=x(n)\hat{\theta} = x_{(n)}, we need to find its expected value, E[x(n)]E[x_{(n)}].

Finding the Expected Value of x(n)x_{(n)}

First, we need the cumulative distribution function (CDF) of x(n)x_{(n)}. The CDF of x(n)x_{(n)}, denoted as Fx(n)(x)F_{x_{(n)}}(x), is the probability that the maximum of the sample is less than or equal to xx:

Fx(n)(x)=P(x(n)x)F_{x_{(n)}}(x) = P(x_{(n)} \leq x)

Since x(n)x_{(n)} is the maximum of the sample, x(n)xx_{(n)} \leq x if and only if all the sample values are less than or equal to xx. Thus,

Fx(n)(x)=P(x1x,x2x,...,xnx)F_{x_{(n)}}(x) = P(x_1 \leq x, x_2 \leq x, ..., x_n \leq x)

Since the xix_i are independent and identically distributed (i.i.d.), we have:

Fx(n)(x)=[P(x1x)]nF_{x_{(n)}}(x) = [P(x_1 \leq x)]^n

For the uniform distribution on (0,θ)(0, \theta), the CDF of a single observation xix_i is:

Fxi(x)={0if x0xθif 0<x<θ1if xθF_{x_i}(x) = \begin{cases} 0 & \text{if } x \leq 0 \\ \frac{x}{\theta} & \text{if } 0 < x < \theta \\ 1 & \text{if } x \geq \theta \end{cases}

Therefore, the CDF of x(n)x_{(n)} is:

Fx(n)(x)={0if x0(xθ)nif 0<x<θ1if xθF_{x_{(n)}}(x) = \begin{cases} 0 & \text{if } x \leq 0 \\ (\frac{x}{\theta})^n & \text{if } 0 < x < \theta \\ 1 & \text{if } x \geq \theta \end{cases}

Next, we find the probability density function (PDF) of x(n)x_{(n)} by differentiating the CDF with respect to xx:

fx(n)(x)=ddxFx(n)(x)={nxn1θnif 0<x<θ0otherwisef_{x_{(n)}}(x) = \frac{d}{dx} F_{x_{(n)}}(x) = \begin{cases} \frac{n x^{n-1}}{\theta^n} & \text{if } 0 < x < \theta \\ 0 & \text{otherwise} \end{cases}

Now we can compute the expected value of x(n)x_{(n)}:

E[x(n)]=xfx(n)(x)dx=0θxnxn1θndx=nθn0θxndxE[x_{(n)}] = \int_{-\infty}^{\infty} x f_{x_{(n)}}(x) dx = \int_{0}^{\theta} x \frac{n x^{n-1}}{\theta^n} dx = \frac{n}{\theta^n} \int_{0}^{\theta} x^n dx

E[x(n)]=nθn[xn+1n+1]0θ=nθnθn+1n+1=nn+1θE[x_{(n)}] = \frac{n}{\theta^n} \left[\frac{x^{n+1}}{n+1}\right]_0^{\theta} = \frac{n}{\theta^n} \frac{\theta^{n+1}}{n+1} = \frac{n}{n+1} \theta

Calculating the Bias

Now we can calculate the bias:

Bias(θ^)=E[x(n)]θ=nn+1θθ=θn+1Bias(\hat{\theta}) = E[x_{(n)}] - \theta = \frac{n}{n+1} \theta - \theta = -\frac{\theta}{n+1}

The bias is negative, which means that on average, the estimator underestimates the true value of θ\theta. The magnitude of the bias decreases as the sample size nn increases, which is a desirable property.

Variance of the Estimator

The variance of an estimator measures its statistical variability or dispersion. A lower variance indicates that the estimator's values are more tightly clustered around its expected value. To find the variance of θ^=x(n)\hat{\theta} = x_{(n)}, we use the formula:

Var(θ^)=E[θ^2](E[θ^])2Var(\hat{\theta}) = E[\hat{\theta}^2] - (E[\hat{\theta}])^2

We already know E[θ^]=E[x(n)]=nn+1θE[\hat{\theta}] = E[x_{(n)}] = \frac{n}{n+1} \theta. Now we need to find E[x(n)2]E[x_{(n)}^2].

Finding E[x(n)2]E[x_{(n)}^2]

E[x(n)2]=x2fx(n)(x)dx=0θx2nxn1θndx=nθn0θxn+1dxE[x_{(n)}^2] = \int_{-\infty}^{\infty} x^2 f_{x_{(n)}}(x) dx = \int_{0}^{\theta} x^2 \frac{n x^{n-1}}{\theta^n} dx = \frac{n}{\theta^n} \int_{0}^{\theta} x^{n+1} dx

E[x(n)2]=nθn[xn+2n+2]0θ=nθnθn+2n+2=nn+2θ2E[x_{(n)}^2] = \frac{n}{\theta^n} \left[\frac{x^{n+2}}{n+2}\right]_0^{\theta} = \frac{n}{\theta^n} \frac{\theta^{n+2}}{n+2} = \frac{n}{n+2} \theta^2

Calculating the Variance

Now we can calculate the variance:

Var(θ^)=E[x(n)2](E[x(n)])2=nn+2θ2(nn+1θ)2Var(\hat{\theta}) = E[x_{(n)}^2] - (E[x_{(n)}])^2 = \frac{n}{n+2} \theta^2 - \left(\frac{n}{n+1} \theta\right)^2

Var(θ^)=θ2[nn+2n2(n+1)2]=θ2[n(n+1)2n2(n+2)(n+2)(n+1)2]Var(\hat{\theta}) = \theta^2 \left[\frac{n}{n+2} - \frac{n^2}{(n+1)^2}\right] = \theta^2 \left[\frac{n(n+1)^2 - n^2(n+2)}{(n+2)(n+1)^2}\right]

Var(θ^)=θ2n(n2+2n+1)n32n2(n+2)(n+1)2=θ2n3+2n2+nn32n2(n+2)(n+1)2Var(\hat{\theta}) = \theta^2 \frac{n(n^2 + 2n + 1) - n^3 - 2n^2}{(n+2)(n+1)^2} = \theta^2 \frac{n^3 + 2n^2 + n - n^3 - 2n^2}{(n+2)(n+1)^2}

Var(θ^)=nθ2(n+2)(n+1)2Var(\hat{\theta}) = \frac{n \theta^2}{(n+2)(n+1)^2}

The variance decreases as the sample size nn increases, which is also a desirable property for an estimator.

Consistency of the Estimator

Consistency is a property that describes the behavior of an estimator as the sample size approaches infinity. An estimator is consistent if it converges in probability to the true value of the parameter. In other words, as the sample size grows, the estimator becomes more and more likely to be close to the true value.

To show that θ^=x(n)\hat{\theta} = x_{(n)} is a consistent estimator for θ\theta, we need to show that for any ϵ>0\epsilon > 0:

limnP(θ^θ>ϵ)=0\lim_{n \to \infty} P(|\hat{\theta} - \theta| > \epsilon) = 0

Equivalently, we can show that:

limnP(x(n)θ<ϵ)=1\lim_{n \to \infty} P(|x_{(n)} - \theta| < \epsilon) = 1

Let's rewrite the probability:

P(x(n)θ<ϵ)=P(θϵ<x(n)<θ+ϵ)P(|x_{(n)} - \theta| < \epsilon) = P(\theta - \epsilon < x_{(n)} < \theta + \epsilon)

Since x(n)x_{(n)} cannot be greater than θ\theta, we only need to consider the lower bound:

P(θϵ<x(n)<θ)=P(x(n)>θϵ)P(\theta - \epsilon < x_{(n)} < \theta) = P(x_{(n)} > \theta - \epsilon)

This is equivalent to saying that at least one of the sample values xix_i is greater than θϵ\theta - \epsilon. It's easier to consider the complement: the probability that all xix_i are less than or equal to θϵ\theta - \epsilon:

P(x(n)>θϵ)=1P(x1θϵ,...,xnθϵ)P(x_{(n)} > \theta - \epsilon) = 1 - P(x_1 \leq \theta - \epsilon, ..., x_n \leq \theta - \epsilon)

P(x(n)>θϵ)=1[P(x1θϵ)]nP(x_{(n)} > \theta - \epsilon) = 1 - [P(x_1 \leq \theta - \epsilon)]^n

Using the CDF of the uniform distribution:

P(x1θϵ)=θϵθ=1ϵθP(x_1 \leq \theta - \epsilon) = \frac{\theta - \epsilon}{\theta} = 1 - \frac{\epsilon}{\theta}

provided that θϵ>0\theta - \epsilon > 0 (i.e., ϵ<θ\epsilon < \theta). Then,

P(x(n)>θϵ)=1(1ϵθ)nP(x_{(n)} > \theta - \epsilon) = 1 - \left(1 - \frac{\epsilon}{\theta}\right)^n

Now, take the limit as nn approaches infinity:

limnP(x(n)>θϵ)=limn[1(1ϵθ)n]\lim_{n \to \infty} P(x_{(n)} > \theta - \epsilon) = \lim_{n \to \infty} \left[1 - \left(1 - \frac{\epsilon}{\theta}\right)^n\right]

Since 0<1ϵθ<10 < 1 - \frac{\epsilon}{\theta} < 1, we have:

limn(1ϵθ)n=0\lim_{n \to \infty} \left(1 - \frac{\epsilon}{\theta}\right)^n = 0

Therefore,

limnP(x(n)>θϵ)=10=1\lim_{n \to \infty} P(x_{(n)} > \theta - \epsilon) = 1 - 0 = 1

This shows that θ^=x(n)\hat{\theta} = x_{(n)} is a consistent estimator for θ\theta.

Summary of Estimator Properties

  • Bias: The estimator is biased, with Bias(θ^)=θn+1Bias(\hat{\theta}) = -\frac{\theta}{n+1}. The bias is negative, indicating underestimation, but it decreases as nn increases.
  • Variance: The variance is Var(θ^)=nθ2(n+2)(n+1)2Var(\hat{\theta}) = \frac{n \theta^2}{(n+2)(n+1)^2}. The variance also decreases as nn increases.
  • Consistency: The estimator is consistent, meaning that it converges in probability to the true value of θ\theta as nn approaches infinity.

These properties provide a comprehensive understanding of the behavior of the MLE for θ\theta in the uniform distribution. While the estimator is biased, it is consistent, and its bias and variance decrease with increasing sample size. This information is crucial for making informed decisions about the accuracy and reliability of the estimate in practical applications. In the next section, we will discuss bias correction techniques to improve the estimator's performance.

Bias Correction

As we have established, the Maximum Likelihood Estimator (MLE) for θ\theta in the uniform distribution, θ^=x(n)\hat{\theta} = x_{(n)}, is biased. Specifically, the bias is given by:

Bias(θ^)=E[θ^]θ=θn+1Bias(\hat{\theta}) = E[\hat{\theta}] - \theta = -\frac{\theta}{n+1}

The fact that the MLE underestimates θ\theta on average motivates us to consider bias correction techniques. Bias correction aims to adjust the estimator so that its expected value is closer to the true parameter value. In this section, we will derive a bias-corrected estimator for θ\theta.

Deriving the Bias-Corrected Estimator

To correct for the bias, we want to find an estimator θ^corrected\hat{\theta}_{corrected} such that:

E[θ^corrected]=θE[\hat{\theta}_{corrected}] = \theta

Let's express the corrected estimator as a linear function of the original estimator:

θ^corrected=cθ^=cx(n)\hat{\theta}_{corrected} = c \hat{\theta} = c x_{(n)}

where cc is a constant that we need to determine. Now, we take the expected value:

E[θ^corrected]=E[cx(n)]=cE[x(n)]=cnn+1θE[\hat{\theta}_{corrected}] = E[c x_{(n)}] = c E[x_{(n)}] = c \frac{n}{n+1} \theta

We want this expected value to be equal to θ\theta, so we set:

cnn+1θ=θc \frac{n}{n+1} \theta = \theta

Solving for cc:

c=n+1nc = \frac{n+1}{n}

Thus, the bias-corrected estimator is:

θ^corrected=n+1nx(n)=n+1nmax(x1,x2,...,xn)\hat{\theta}_{corrected} = \frac{n+1}{n} x_{(n)} = \frac{n+1}{n} \max(x_1, x_2, ..., x_n)

This estimator is unbiased because its expected value is equal to the true parameter value θ\theta.

Properties of the Bias-Corrected Estimator

Now that we have the bias-corrected estimator, it's important to examine its properties. We know it is unbiased by construction, but let's consider its variance.

Variance of the Bias-Corrected Estimator

The variance of the bias-corrected estimator is:

Var(\hat{\theta}_{corrected}) = Var\left(\frac{n+1}{n} x_{(n)}\right) = \left(\frac{n+1}{n}\right)^2 Var(x_{(n)})$ ## Bias-Corrected Estimator Conclusion The **bias-corrected estimator** not only provides an unbiased estimate of $\theta$ but also demonstrates the power of refining statistical estimates through mathematical adjustments. This meticulous approach to estimation is what makes statistical inference a powerful tool in various fields, from scientific research to data analysis in business and economics. # Conclusion In this comprehensive exploration, we've delved into the intricacies of estimating the parameter $\theta$ for a uniform distribution given a random sample $x_1, ..., x_n$. We began by understanding the fundamental principles of Maximum Likelihood Estimation (MLE), a cornerstone of statistical inference. The uniform distribution, characterized by its equal probability across a defined interval, presents a unique challenge and opportunity for parameter estimation. ## MLE Estimation Our journey through the **MLE** process led us to a seemingly simple yet profound result: the MLE estimator for $\theta$ is the sample maximum, denoted as $x_{(n)}$. This means that the best estimate for the upper bound of the uniform distribution, based on the observed data, is the largest value in the sample. The intuition behind this result is clear – the parameter $\theta$ cannot be smaller than the largest observation, and the likelihood decreases as $\theta$ increases beyond this point. However, this estimator, while intuitive, is not without its flaws. We discovered that the MLE estimator is biased, tending to underestimate the true value of $\theta$. The bias, quantified as $-\frac{\theta}{n+1}$, reveals a systematic underestimation that diminishes as the sample size grows. ## Bias Properties The **bias** of an estimator is a critical property that reflects its tendency to deviate from the true parameter value. Recognizing the bias in our MLE estimator prompted us to investigate techniques for bias correction. We rigorously derived a bias-corrected estimator by scaling the original MLE estimator. This corrected estimator, given by $\hat{\theta}_{corrected} = \frac{n+1}{n} x_{(n)}$, ensures that, on average, our estimate aligns with the true value of $\theta$. The journey did not end with bias correction. We extended our analysis to evaluate the variance of both the original and bias-corrected estimators. Variance measures the spread or dispersion of the estimator's values, indicating its precision. A lower variance signifies a more precise estimator. Our calculations revealed that the bias-corrected estimator, while unbiased, exhibits a slightly higher variance compared to the original MLE estimator. This highlights a common trade-off in statistical estimation: reducing bias can sometimes increase variance, and vice versa. ## Consistency Discussion Another key property we examined was **consistency**. A consistent estimator is one that converges to the true parameter value as the sample size increases. We demonstrated that the original MLE estimator, despite its bias, is consistent. This means that with a sufficiently large sample, the estimator will provide a reliable estimate of $\theta$. The bias-corrected estimator, being a scaled version of the consistent MLE estimator, is also consistent. In conclusion, this exploration has provided a comprehensive understanding of parameter estimation for the uniform distribution. We have navigated the landscape of MLE, bias, variance, and consistency, uncovering the nuances of statistical estimation. The insights gained here are not only valuable for this specific problem but also serve as a foundation for tackling more complex estimation challenges in statistics and data science.