The Personal Site of Lalo Morales


Foundations for Building Large Language Models: A Beginner-Friendly Roadmap

LLM foundations w andrej karpathy

Inspired by Andrej Karpathy’s curriculum and journey, this report guides motivated beginners through the essential foundations needed to understand and build large language models (LLMs). We’ll cover fundamental math and CS knowledge, core machine learning concepts, modern NLP architectures like Transformers, and effective learning strategies. The aim is to provide a clear, structured study plan – with simple explanations, code examples, diagrams, and references to further resources – serving as a long-term roadmap for your LLM ambitions.

I. Foundational Knowledge

Overview: Modern AI and LLMs sit atop a bedrock of mathematics and computer science fundamentals. Developing “from scratch” implementations (a hallmark of Karpathy’s teaching (Neural Networks: Zero To Hero)) requires comfort with linear algebra, calculus, probability, and more, as well as an understanding of how computers store and process data. In this section, we outline the key foundational topics and why they matter.

Mathematics Essentials

Linear Algebra: At the heart of neural networks are vectors and matrices – for example, model weights can be seen as matrices transforming input vectors into output vectors. Linear algebra is crucial for understanding many ML algorithms (especially deep learning) (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.). Key concepts include: vectors and matrices (which represent data and transformations in models), linear transformations (operations like rotations or scaling of data), eigenvalues and eigenvectors (which, for a matrix, indicate fundamental modes of transformation), and decompositions like SVD (Singular Value Decomposition) and PCA (Principal Component Analysis) (used for dimensionality reduction and understanding data variance). For instance, PCA uses linear algebra to find the principal components (eigenvectors of the data’s covariance matrix) that explain the most variance in high-dimensional data. Example: computing eigenvalues of a matrix with NumPy:

This code finds eigenvalues of matrix A. The output shows the eigenvalues (in this case, 5.0 and 2.0), confirming that one eigenvector direction is “stretchier” (associated with 5) than the other (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.). Understanding such linear algebra results helps in grasping concepts like PCA or how a neural network layer might scale certain directions of input data more than others.

Calculus (Multivariable Calculus): Training neural networks involves optimizing a loss function with respect to many parameters. Calculus provides the language of gradients – vectors of partial derivatives – that tell us how to update parameters to reduce loss. Key topics include derivatives (rate of change of a function), gradient vectors (for multi-variable functions, indicating the direction of steepest ascent/descent), and integrals (useful for understanding continuous distributions, areas under curves, etc.). In deep learning, backpropagation is essentially repeated application of the chain rule from calculus to compute gradients of the loss with respect to each parameter. A solid grasp of calculus is needed to understand how model parameters are tuned during training (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.). For example, one might use calculus to derive that the gradient of the mean squared error (MSE) loss with respect to a linear model’s weight is proportional to the error, which informs how we adjust the weight.

Probability and Statistics: Machine learning learns from data, so understanding uncertainty and data distributions is vital. Probability theory covers random variables and distributions (e.g. normal distribution, which often underlies weight initialization in neural nets), while statistics covers estimation and inference (like estimating the mean of data or testing hypotheses). Concepts like Bayes’ Theorem form the basis of Bayesian machine learning and help in understanding how to update beliefs with new evidence. In training LLMs, we often treat the next-word prediction as a probability distribution (the model outputs a probability for each possible token). Knowledge of expected values, variance, and covariance helps in understanding algorithms and evaluating model performance (e.g. variance in model predictions). Statistics also informs evaluation metrics and significance testing. In summary, probability & statistics are crucial for understanding how models learn from data and how to interpret their predictions (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.). As an example, understanding the cross-entropy loss used in classification requires interpreting it as the negative log-likelihood of the correct class under the model’s predicted probability distribution – a concept rooted in information theory and probability.

Information Theory: Closely linked to probability, information theory provides measures of uncertainty and information content. An important concept is entropy, which quantifies the uncertainty (or surprise) in a random variable’s outcomes. High entropy means outcomes are very unpredictable (requiring more information to describe) (Information Theory in Machine Learning | GeeksforGeeks). For instance, a fair 6-sided die has higher entropy than a fair coin toss. In ML, entropy shows up in the loss functions: the cross-entropy loss used to train classifiers and language models is derived from entropy. Lower cross-entropy means the model’s predicted distribution is closer to the true distribution of data. Mutual information measures how much knowing one variable reduces uncertainty about another – useful for feature selection (which features give most info about the label) (Information Theory in Machine Learning | GeeksforGeeks). Information theory also underpins information gain used in decision tree splits and concepts like perplexity (the exponentiated entropy, used to evaluate language models – a lower perplexity means the model is less “surprised” by the test data). In summary, information theory gives a formal way to think about learning as reducing uncertainty. For example, training a language model reduces its entropy over text, meaning it becomes more confident and accurate in predicting words (Information Theory in Machine Learning | GeeksforGeeks).

Computer Science Fundamentals

Data Structures and Algorithms: Building efficient ML systems requires knowledge of how to store and manipulate data. Data structures (like arrays, lists, trees, hash maps) organize data for efficient access and updates. Algorithms (like sorting, searching, graph traversal) provide the steps to solve computational problems. ML implementations often rely on optimized algorithms – for example, computing the nearest neighbors in a dataset, balancing a decision tree, or shuffling data for training. Strong CS fundamentals help you reason about the computational complexity of training algorithms (big-O notation) and to write code that scales to large data. In fact, knowing algorithms is important not just for coding interviews but for ML itself: Many advanced ML techniques (like dynamic programming for the Viterbi algorithm in HMMs, or graph algorithms in message passing networks) come directly from classical algorithms. Moreover, efficient code can make the difference between a training run finishing overnight versus in a week. As one practitioner put it, machine learning and data structures & algorithms go hand-in-hand: ML finds patterns in data, and DSA enables storing and processing that data efficiently (Why DSA Important for Machine Learning?). Data structures help “store data efficiently and write optimized programs” (Why DSA Important for Machine Learning?) – for example, using a hash table to memoize results can speed up repetitive computations in training loops. Many real-world ML projects involve custom preprocessing or postprocessing code where algorithmic thinking is required. In short, a good ML engineer should appreciate that while ML models learn from data, it’s efficient algorithms that feed them that data and deploy them at scale.

Operating Systems (OS) and Computer Architecture: While at first glance training a model seems abstracted away from low-level details, understanding how software and hardware interact is highly beneficial. An OS manages memory, processes, and hardware resources. Training large models (like GPT-style transformers) is resource-intensive – knowledge of OS concepts like memory management (to avoid out-of-memory errors), process scheduling (for multi-GPU training or data loading in parallel), and file systems (for efficient data streaming from disk) helps in optimizing training jobs. For instance, setting up a distributed training job requires understanding how processes coordinate and share data (often via OS primitives or libraries built on them). Knowledge of computer architecture is equally useful: modern ML relies on specialized hardware (GPUs, TPUs) that exploit parallelism. Understanding the basics of CPU vs GPU architecture, memory hierarchies (cache vs RAM vs VRAM), and instruction pipelines can help you write more efficient code (e.g., using vectorized operations and avoiding memory bottlenecks). As an example, matrix multiplications are much faster on GPUs due to their parallel structure – knowing this, you’d aim to offload such computations appropriately. As one expert put it, students need to understand computer architecture to structure a program so that it runs efficiently on real machines (How does understanding computer architecture help a programmer?). Similarly, OS knowledge is “always handy” in ML – it helps you understand “what is happening under the hood” when your code interfaces with hardware ([D] how useful is OS knowledge in AI/ML? : r/MachineLearning). In practice, this might mean knowing how to optimize data loading (e.g., using asynchronous I/O or memory mapping large datasets) or how to avoid CPU-GPU data transfer bottlenecks. In sum, while you don’t need to be an OS developer to do ML, appreciating these systems concepts will make you a much more effective practitioner, especially as you scale up to real-world LLM training which can involve distributed systems, cloud computing, and optimization at the hardware level.

Programming and Software Engineering: (Not explicitly listed in the question, but worth a brief mention) Alongside theory, you’ll need solid programming skills – Python is the de facto language for ML, so being comfortable with Python’s syntax and libraries (NumPy for math, pandas for data manipulation, etc.) is assumed (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.). Good coding practices (modular design, version control, testing) are invaluable when your projects grow in complexity (for example, training an LLM might involve coordinating multiple scripts for data processing, model training, and evaluation). Andrej Karpathy emphasizes “hacker” skills – the ability to quickly prototype and experiment – which come from lots of practical coding experience.

II. Machine Learning & Deep Learning

Overview: With the math and CS toolkit in hand, the next step is understanding how machines learn patterns from data. This involves two layers: Classical Machine Learning algorithms that form the basis of modeling, and Deep Learning techniques that have driven recent breakthroughs in AI (including LLMs). We’ll start with a brief tour of classical ML, then delve into neural networks and modern deep learning architectures. We’ll also cover model training techniques (optimization algorithms, regularization, evaluation) that apply across both paradigms.

Classical Machine Learning

Before diving into deep neural networks, it’s helpful to understand some classical ML algorithms and concepts. These algorithms often provide simpler models that are easier to interpret and are effective on smaller datasets or as building blocks. A few key ones include:

  • Regression: Regression is the task of predicting a continuous quantity from input features (1. What is machine learning?? What does learning imply in the…). The simplest example is linear regression, where we fit a line (or hyperplane in higher dimensions) to data points. For instance, predicting house prices based on size can be done with linear regression by fitting a line $price = w \cdot size + b$. Even though deep learning can also do regression, understanding linear regression is important – it introduces the idea of a loss function (like mean squared error) and how we can solve for model parameters (analytically or via gradient descent). More advanced forms include logistic regression (actually a classification algorithm despite the name, using a logistic function to predict probabilities of classes). Code Example: fitting a simple linear regression using scikit-learn:

Output:

Learned coefficient (slope): 1.9999999999999996  

Learned intercept: 1.0000000000000009  

Prediction for x=6: 12.999999999999996

As expected, the model finds a slope ~2 and intercept ~1, and predicts ~13 for x=6. Understanding this simple case helps in grasping how more complex models learn parameters to fit data.

  • Classification (SVMs, Decision Trees, etc.): Classification is the task of predicting a discrete class label (e.g., spam vs. not-spam email) (What is Classification in AI?). There are several popular classical methods:
    • Support Vector Machines (SVMs): SVMs are supervised learning models that can perform classification (and regression). Intuitively, an SVM finds the optimal separating hyperplane between classes by maximizing the margin (distance) between the nearest points of each class and the decision boundary (What Is Support Vector Machine? | IBM). This results in a robust classifier that generalizes well. SVMs can also use kernel tricks to handle non-linear decision boundaries by implicitly mapping data to higher-dimensional spaces. In practice, SVMs were a dominant method for many problems before deep learning took over; they’re still useful for smaller datasets. Key takeaway from SVM: the concept of a margin and the idea that not all training points matter for defining the decision boundary (only the support vectors do).
    • Decision Trees: A decision tree is a flowchart-like model of decisions. It repeatedly splits the data based on features, forming a tree where each internal node is a binary (or multi-way) test on a feature, each branch is an outcome of the test, and each leaf node represents a predicted class or value (What is a Decision Tree? | IBM). For example, a simple tree might classify if one should play tennis based on weather: first split on “Is it sunny?”, then maybe split on humidity, etc., until reaching a decision (play or not). Trees are easy to interpret (you can trace the path of decisions) and form the basis of powerful ensemble methods like Random Forests. However, single trees can overfit, which is why pruning (cutting back the tree to limit depth or complexity) is used. Decision trees introduce important concepts like information gain (often measured via entropy reduction (What is a Decision Tree? | IBM)) to decide which split is best.
    • Other classical algorithms: There are many more, like k-Nearest Neighbors (which classifies a point based on the majority label of its $k$ nearest points in the training set), Naive Bayes (which applies Bayes’ theorem with an assumption of feature independence), K-means clustering (unsupervised, for grouping data into $k$ clusters), and so on. Each algorithm has its own assumptions and suited scenarios.

Why study these if we have deep learning? Because they build intuition for different ways of learning from data. They are also still used in practice (not every problem needs a deep net; for example, a simple regression or tree might suffice for a small dataset and be more interpretable). Karpathy himself, in teaching, often references simple models as sanity checks or starting baselines (Neural Networks: Zero To Hero). Moreover, ensemble methods like boosting (e.g., XGBoost) which are based on decision trees remain top performers on structured data competitions.

Deep Neural Networks and Transformers

While classical ML is important, the rise of deep learning has revolutionized AI. Deep learning refers to neural networks with multiple layers (“deep” refers to many layers of neurons) that can learn complex representations of data. Karpathy famously taught Stanford’s CS231n on Convolutional Neural Networks; here we’ll summarize the main types of neural networks leading up to transformers:

Neural Network Basics: A neural network is composed of layers of interconnected “neurons”. Each neuron computes a weighted sum of inputs and applies a nonlinear activation function. By stacking layers, neural networks can approximate complex functions. Key components:

  • Feed-Forward Neural Networks (MLPs): These are the simplest deep networks, also called multi-layer perceptrons. They consist of an input layer, one or more hidden layers (each a set of neurons with nonlinear activations like ReLU), and an output layer. Every neuron in layer L connects to every neuron in layer L+1 (hence “fully connected layers”). MLPs can model nonlinear relationships, but for high-dimensional inputs like images or sequences, they don’t scale well without special structure.
  • Activation Functions: Nonlinear functions like ReLU (Rectified Linear Unit), sigmoid, or tanh applied at each neuron. They introduce non-linearity, which is crucial – without them, multiple layers would collapse into an equivalent single linear layer.

Convolutional Neural Networks (CNNs): CNNs are specialized for grid-like data such as images. Instead of fully connecting every input pixel to a neuron, CNNs use convolutional layers that apply a small filter (kernel) that slides across the image, detecting local patterns (like edges, textures) and producing feature maps. This introduces the idea of local receptive fields and weight sharing (the same filter is applied across the image), which drastically reduces parameters and encodes the assumption that local patterns can appear anywhere in the image. After convolutional layers (and pooling layers that downsample resolution), a CNN usually has fully connected layers to produce the final classification. CNNs have been extremely successful for vision tasks – e.g., the AlexNet CNN was a breakthrough on ImageNet classification in 2012. The concept to grasp: CNNs automatically learn hierarchical feature representations (first layers learn edges, later layers learn object parts, etc.). As a beginner, you should understand how a convolution operation works and why it helps in images (translational invariance, etc.). In summary, CNNs process grid-like data (like images) using convolution operations to extract spatial features (Convolutional Neural Network (CNN) in Machine Learning | GeeksforGeeks). Modern CNN architectures (ResNet, etc.) add innovations like residual connections, but all build on this core idea.

Recurrent Neural Networks (RNNs): RNNs are designed for sequential data (like time series or natural language). An RNN processes one element of the sequence at a time while maintaining a hidden state that carries information about previous elements. In effect, RNNs have a “memory” – the output at time t can influence the computation at time t+1. This makes them ideal for language modeling (predicting the next word given previous words) or other sequence tasks, since they naturally encode order and context. A simple RNN has an equation like $h_t = f(W x_t + U h_{t-1})$, where $h_t$ is the hidden state and $x_t$ the input at time $t$. They are called “recurrent” because the hidden state recurs across time steps. However, basic RNNs suffer from difficulties learning long-range dependencies (due to vanishing/exploding gradients over long sequences). This led to LSTM (Long Short-Term Memory) networks and GRUs (Gated Recurrent Units), which are RNN variants with gating mechanisms that better control information flow and can maintain long-term information. The key concept is that RNNs incorporate feedback loops allowing earlier outputs/hidden states to influence later computations, thereby capturing temporal dependencies (Recurrent neural network – Wikipedia). For example, in a sentence, an RNN’s hidden state can carry context from previous words to help predict the next word. Karpathy’s blog “The Unreasonable Effectiveness of RNNs” demonstrated their ability to generate character-level text by learning dependencies in sequences. While RNNs have largely been eclipsed by Transformers for NLP, they’re still worth learning as they introduce the idea of sequence modelling and were state-of-the-art for tasks like language translation and speech recognition for many years.

Transformers and Self-Attention: The Transformer architecture has now become the foundation for large language models. Introduced in the landmark paper “Attention is All You Need” (Vaswani et al. 2017), Transformers dispense with recurrence and instead rely entirely on a mechanism called self-attention to handle sequence data ([PDF] Attention is All you Need – NIPS papers) (Transformers). Below is a diagram of the Transformer’s encoder-decoder architecture (with the encoder on the left and decoder on the right):

(File:Transformer, full architecture.png – Wikimedia Commons) Diagram: The Transformer architecture (encoder on left, decoder on right). Each encoder layer (yellow/orange outline) has self-attention and feed-forward sublayers, and each decoder layer (green outline) has self-attention, cross-attention (attending to encoder outputs), and feed-forward sublayers. This architecture enables modeling of long-range dependencies with ease, as every token can attend to every other token.

How Self-Attention Works: In a nutshell, self-attention allows each position in a sequence to dynamically focus on other positions. For each input token (e.g., a word or subword), the model computes queries, keys, and values (Q, K, V) as vectors. The attention mechanism computes a weighted sum of value vectors for each token, where the weights come from the dot-product similarity between that token’s query and other tokens’ keys. This means if one word is very relevant to another, it will have a high weight (attention) when computing the other’s representation. For example, in the sentence “The cat sat on the mat,” a self-attention head might learn to link “cat” and “mat” (perhaps because they rhyme or are related in context) by giving them higher attention to each other’s representations. The Transformer uses multiple attention heads (“multi-head attention”) to capture different types of relationships. By stacking multiple self-attention layers (with residual connections and layer normalization for stability), the encoder can produce rich contextualized representations for each token considering the whole sequence (Transformers).

Transformer Encoder and Decoder: The original Transformer has an encoder-decoder structure. The encoder reads the input sequence (e.g., an English sentence in a translation task) and produces a sequence of continuous representations. The decoder then takes those representations and generates an output sequence (e.g., the French translation), one token at a time, using both self-attention (to consider what it has generated so far) and cross-attention to the encoder’s outputs (so it can condition on the input sentence). The decoder is often trained with a technique called teacher forcing where it is given the true previous token during training, and at inference it uses its own generated tokens. The key innovation of the Transformer is that it forgoes RNNs’ step-by-step recurrence and instead processes sequences in parallel, attending to all positions at once in each layer. This makes it much more efficient to train on GPUs (where parallel computation is crucial) and it captures long-range dependencies better (since even distant tokens can be attended to directly without having to pass through many intermediate steps).

GPT, BERT, T5 (Examples of Transformer-based models): Once Transformers appeared, various researchers adapted the architecture for different purposes:

  • GPT (Generative Pre-trained Transformer): This is a decoder-only Transformer (essentially just the Transformer decoder stack, without an encoder) trained in an unsupervised way to predict the next token in large-scale text (Transformers – Hugging Face). OpenAI’s GPT-1, GPT-2, and GPT-3 demonstrated the power of this approach – after training on a huge corpus, the model can generate coherent text and be fine-tuned or prompted for various tasks. GPT is unidirectional (it only attentively looks at past tokens when generating the next token). It’s “pre-trained” on generic internet text, then can be fine-tuned or prompted for downstream tasks. The takeaway: GPT is an autoregressive language model that can generate text; it treats language modeling as the central training objective. This simple setup, when scaled up, produces surprisingly powerful capabilities in the model.
  • BERT (Bidirectional Encoder Representations from Transformers): BERT, developed by Google ((PDF) INISTA Log and Execution Trace Analytics System | Murat C …) (What is the BERT language model? | Definition from TechTarget), is an encoder-only Transformer. It’s trained with a different objective: masked language modeling (randomly masking some words in a sentence and training the model to predict them) and next sentence prediction (predict if one sentence follows another). BERT’s architecture allows it to attend bidirectionally – meaning a word’s representation can be influenced by both its left and right context (something GPT’s one-directional approach doesn’t allow for that word prediction task). After pre-training on enormous text (like Wikipedia), BERT produces powerful contextual embeddings that can be fine-tuned for tasks like question answering, sentiment analysis, etc. Essentially, BERT acts as a deep understanding model: it produces a rich representation of text that captures context from both sides. For example, BERT will represent the word “bank” differently in “river bank” vs “bank account” by looking at surrounding words. BERT “uses the transformer’s bidirectional attention to understand ambiguous language in text by using surrounding text to establish context” (What is the BERT language model? | Definition from TechTarget) (What is the BERT language model? | Definition from TechTarget). It revolutionized NLP by enabling transfer learning – one can fine-tune BERT on a small dataset for a specific task and get state-of-the-art results because BERT already learned a lot of general language patterns.
  • T5 (Text-To-Text Transfer Transformer): T5, from Google, takes the idea of “everything is a text-to-text problem” (google-t5/t5-base – Hugging Face). It’s a sequence-to-sequence (encoder-decoder) Transformer pre-trained on a multitask mixture, all cast into a text-to-text format. For example, translating English to French, summarizing an article, or answering a question are all trained by feeding in some text prompt and training the model to output the target text. T5’s contribution was to unify NLP tasks under the same training framework and demonstrate excellent performance by scaling up. In practice, using T5 means you can fine-tune it with a prefix indicating the task (e.g., “translate English to French: ”) and it will generate the output. T5 showed that with sufficient scale, a single model can learn to perform a wide variety of tasks by casting them as text generation problems (google-t5/t5-base – Hugging Face). It also introduced the concept of “Colossal Clean Crawled Corpus (C4)” as a large clean pre-training dataset.

Why Transformers matter for LLMs: Large Language Models like GPT-3 are essentially very large Transformers (with dozens of layers, thousands of attention heads, and billions of parameters) trained on very large text corpora. They demonstrate emergent capabilities (like few-shot learning, where the model can perform a task given only a few examples in the prompt without gradient updates). The field has observed scaling laws (discussed below) that as you increase the model size, data size, and compute, the performance keeps improving (Issue #170 – Data and Parameter Scaling Laws for Neural Machine Translation). Transformers have become the architecture of choice because they scale efficiently and learn very expressive representations of sequences. Thus, to build or understand an LLM, one must understand Transformers and attention.

Optimization Algorithms: Regardless of model type (whether a simple regression or a deep network), training involves optimizing a loss function. The most common approach is gradient descent – iteratively adjusting parameters in the opposite direction of the gradient of the loss. In practice, we use Stochastic Gradient Descent (SGD), which means we use a batch of data at a time (not the entire dataset) to compute an approximate gradient, which introduces some noise but is much faster and can generalize better. There are many variants/improvements:

  • Momentum: Accelerates SGD by accumulating an exponentially decaying moving average of past gradients (like giving a velocity to the parameter updates).
  • Adam: A very popular optimizer that combines ideas of momentum and adaptive learning rates for each parameter (Neural Networks: Zero To Hero). Adam keeps track of both the first moment (average) and second moment (uncentered variance) of gradients and scales the learning rate for each parameter inversely to the variance of its gradients. This often leads to faster convergence and is almost a default choice in training neural nets.
  • Learning Rate Schedules: Tuning the learning rate (how big a step we take each update) over time can significantly affect training. Common schedules include decaying the learning rate over epochs or using more advanced schedules (cosine decay, cyclic learning rates, etc.). The learning rate is arguably the most important hyperparameter to tune (Neural Networks: Zero To Hero) – too high and training diverges, too low and it converges slowly or gets stuck.

Code Example (Gradient Descent with PyTorch): To illustrate a basic training loop, consider we want to train a simple one-layer neural network (essentially linear regression) using PyTorch to fit the relationship $y = 2x + 1$. We can use PyTorch’s autograd and optim modules:

After training, the model’s learned weight and bias will be very close to 2 and 1, and the prediction for x=6 will be ~13.0, as expected. This code demonstrates the typical training loop: forward -> compute loss -> backward -> update. In complex models, frameworks like PyTorch handle the gradient calculation, but understanding what happens under the hood (as Karpathy explains in his “micrograd” backpropagation video (Neural Networks: Zero To Hero)) is invaluable.

Regularization: To ensure that models generalize and don’t just memorize the training data (overfitting), we use regularization techniques:

  • L1/L2 Regularization: Adding a penalty term to the loss for large weights (L2 is sum of squares of weights, L1 is sum of absolute values). This encourages the model to keep weights small (which often implies a simpler model). L1 can also drive some weights to exactly 0, effectively performing feature selection.
  • Dropout: A very popular technique for neural nets where, during training, you randomly “drop out” (set to zero) a fraction of the neurons in a layer (Neural Networks: Zero To Hero). This forces the network to not rely too much on any single neuron and leads to more robust representations. At test time, no dropout is applied, but effectively the network’s output is an average of many “thinned” networks from training. Karpathy in his series notes dropout as one of the key innovations that made training deep nets feasible (Neural Networks: Zero To Hero).
  • Batch Normalization: Although originally introduced to help with training convergence by normalizing layer inputs, batch norm also has a side effect of slight regularization. By adding a bit of noise (due to mini-batch fluctuations) and scaling, it can improve generalization. In Karpathy’s videos, batch norm is highlighted for stabilizing deep networks’ training (Neural Networks: Zero To Hero).
  • Early Stopping: Simply stop training when validation performance stops improving, to avoid overfitting the training set.

Loss Functions and Evaluation Metrics: We train models by minimizing a loss function (also called cost function). Selecting the right loss is important:

  • For regression, common losses are Mean Squared Error (MSE) or Mean Absolute Error.
  • For classification, Cross-Entropy Loss (also known as negative log-likelihood) is standard. It’s connected to information theory: it measures the difference between the predicted probability distribution and the true distribution (which is a one-hot vector for the correct class). Minimizing cross-entropy is equivalent to maximizing the likelihood of the data under the model.
  • For specific tasks, there are task-specific losses (e.g., BLEU score for translation, though that’s usually used as an evaluation metric rather than directly optimized).

After training, we evaluate models with appropriate metrics:

  • Classification: accuracy (percent of correct predictions) is simple, but for imbalanced data we look at precision, recall, F1-score, etc. If it’s multi-class, we might use confusion matrices or AUC-ROC for binary classification probability quality.
  • Regression: metrics like R^2 (coefficient of determination), or just reporting the MSE/MAE on a test set.
  • Language models: perplexity (which is $2^{\text{cross-entropy}}$) is commonly reported – it’s the effective average branching factor the model predicts, lower is better. For text generation or translation, BLEU or ROUGE (which compare overlaps with reference texts) are used.

It’s important to use a separate validation set (and ultimately a held-out test set) to evaluate these metrics, to ensure the model generalizes beyond the training data. Karpathy often emphasizes looking not just at final metrics but also at loss curves and potentially even examples of model outputs to truly understand performance (Neural Networks: Zero To Hero) (Neural Networks: Zero To Hero).

By mastering these machine learning fundamentals – from linear regression and SVMs to neural network training and regularization – you build the intuition needed to tackle large-scale problems. Moreover, these concepts often resurface in LLM building: for instance, understanding overfitting and regularization is crucial when fine-tuning an LLM on a small dataset, and knowing about optimization helps when you have to adjust training schedules for large models.

III. NLP and Large Language Models (LLMs)

Overview: With the general ML toolkit covered, we now focus on the specific domain of Natural Language Processing (NLP), which is central to LLMs. We will cover how text data is handled (tokenization, embeddings), the breakthrough Transformer architecture (as introduced above) and specific LLM-related techniques and considerations: scaling laws that guide the growth of model size and dataset, how reinforcement learning from human feedback is used to align LLMs with human preferences, and the emerging skill of prompt engineering to effectively use these models.

Fundamentals of Language Processing

Tokenization: Computers deal with numbers, not words. Tokenization is the process of converting text into tokens (basic units) that the model can understand. Early approaches used words as tokens (a “word-level” tokenizer), but this can be problematic for large vocabularies and misspellings or rare words. Modern approaches often use subword tokenization – for example, Byte Pair Encoding (BPE) or WordPiece – which breaks text into pieces like “unbelievable” -> “un”, “##believ”, “##able”. Using subwords means the model has a manageable vocabulary (e.g., 30k tokens) and can compose words from subword pieces, handling unknown words gracefully. Each token is typically mapped to an integer ID via a vocabulary. For instance, GPT-3 uses the byte-level BPE which can even split a word into characters if it’s very rare, ensuring no out-of-vocabulary tokens. Understanding tokenization is important because it affects how text is represented internally. When building an LLM, one must choose a tokenization scheme that balances vocabulary size and text length (more tokens if using smaller pieces). As a beginner, you could experiment with tokenizing a sentence using a library (like Hugging Face’s transformers library) to see how words break down.

Embeddings: After tokenization, each token needs to be represented as a vector for the model to process. Embeddings are dense vector representations of tokens in a continuous vector space. The idea is to capture semantic similarity – for example, in a good embedding space, synonyms will have similar vectors. In an LLM (or any language model), the first layer is often an embedding lookup: it has an embedding matrix (size vocab_size x embedding_dim) and each token’s ID is used to index a row (the embedding vector). These vectors are learned during training (or one can use pre-trained embeddings like GloVe or fastText in some cases). In Transformers, we also add positional embeddings to encode the position of each token in the sequence, since self-attention alone is order-agnostic (it sees the set of tokens without inherent order). Positional embeddings can be simple (like a sinusoidal pattern as in the original Transformer) or learned vectors. The end result is that each token + position is mapped to an initial vector, which is then processed by the Transformer layers. You can think of embeddings as the model’s “language understanding prior” – early in training they’ll be random, but eventually the embedding for “cat” might end up near “dog” in the vector space, reflecting that the model learned both are animals, etc. A fun project is to train a simple word2vec model (which uses a shallow neural network to learn word embeddings) – it gives insight into how embedding spaces can capture analogies (e.g., the famous example: King – Man + Woman ≈ Queen in the embedding space).

Attention Mechanism Recap: We covered Transformers’ self-attention above. To reinforce: attention is the idea of focusing on parts of the input when producing each part of the output. An LLM with self-attention can, for example, understand that in the sentence “Alice told Bob she would visit him tomorrow,” the word “she” likely refers to Alice (the model can attend from “she” to “Alice” to interpret it) and “him” refers to Bob (Transformers). These pronoun resolution cases are handled elegantly by attention heads. As you work with LLMs, you’ll encounter terms like “attention head”, “query-key-value”, “softmax” – it’s worth demystifying these by maybe writing a small example. For instance, take a short sequence of tokens and manually compute a single-head attention score by creating random query/key/value vectors; this will solidify how the weighted average is computed.

Transformer Architectures: GPT vs BERT vs T5

Let’s summarize the differences of the prominent Transformer-based models, as this is often confusing for newcomers:

  • GPT (Generative Pre-trained Transformer): Architecture: Decoder-only Transformer. It’s trained to predict the next token (so it is an autoregressive language model). Properties: Unidirectional (at generation time it can only look at past context; during training it masks future tokens to prevent cheating). Usage: Great for text generation. By “pre-training” on a large corpus, GPT learns grammar, facts, and some reasoning ability. Later “GPT” models (GPT-2, GPT-3) are essentially just scaled-up versions with more layers, larger embeddings, and trained on more data. They are usually fine-tuned or prompted for tasks (GPT-3, for example, is typically used via prompt engineering since fine-tuning 175B parameters is non-trivial for most). Because GPT is generative, it’s the model behind applications like ChatGPT (with additional fine-tuning). From a learning perspective, training a GPT model from scratch involves feeding tons of text and using cross-entropy loss to make the model’s predicted distribution match the actual next-token distribution. This simple objective belies the power that emerges when the model is huge and trained on virtually the entire internet.
  • BERT (Bidirectional Encoder representations from Transformers): Architecture: Encoder-only Transformer. It uses bidirectional self-attention (every token attends to every other in the input). Training: Not exactly next-word prediction; instead, BERT uses masked language modeling (MLM). For example, “The [MASK] sat on the mat” -> the model should predict “cat”. Because of the bidirectional context, BERT does extremely well on understanding tasks where context on both sides is important. It also had the next sentence prediction (NSP) objective to help it understand sentence relationships. Usage: BERT is typically not used to generate text (you can’t easily sample from it since it’s not a directed generative model). Instead, after pre-training, you add a small layer on top for a specific task and fine-tune BERT. E.g., add a classification layer for sentiment analysis. BERT’s legacy is that it provided a powerful feature extractor for language. Many derivatives like RoBERTa (which dropped the NSP objective and trained longer with more data) improved on BERT, and other models like DistilBERT showed you can compress it. But the core idea remains: bidirectional context for deep understanding (What is the BERT language model? | Definition from TechTarget) (What is the BERT language model? | Definition from TechTarget). BERT’s success means that as an LLM engineer, you might use BERT-like models when you need a strong encoder (for instance, sentence embeddings or as part of a question-answering system where you encode documents and queries).
  • T5 (Text-to-Text Transfer Transformer): Architecture: Encoder-decoder Transformer (like the original sequence-to-sequence model). Training: Multi-task text-to-text. For example, one training example might be: input: “translate English to French: The cat sat on the mat.” output: “Le chat s’est assis sur le tapis.”; another might be input: “summarize: Alice had a cat. The cat sat on the mat.” output: “Alice’s cat sat on the mat.” etc. By unifying tasks, T5 learns a very flexible conditioning – the prefix of the input (before the colon) tells it what task to do. Usage: You can fine-tune T5 on a specific task by continuing training on that task’s examples (often done in the text-to-text format). Or even do zero-shot if the task was sufficiently covered by the multitask mixture. T5 showed the benefit of a unified text-to-text framework where everything (translation, Q&A, classification – framed as “generate the label text”, etc.) is under one roof (google-t5/t5-base – Hugging Face). For someone building LLMs, T5 provides a template for how to set up multi-task training and how to leverage an encoder-decoder model for both understanding and generation. Note that many LLMs nowadays (like GPT-3) are decoder-only, but encoder-decoder models are still used especially in tasks like translation or summarization (e.g., Google’s Pegasus for summarization is encoder-decoder).

In practice, differences between these architectures also influence prompt engineering and fine-tuning approaches. For example, GPT models you can prompt with a few examples in plain text, whereas BERT requires a task-specific head (or using its masked prediction mechanism in creative ways, as was done in some early “prompting” research for BERT).

Scaling Laws: A striking discovery in recent years (by OpenAI’s Kaplan et al. 2020) is that language model performance improves in a predictable power-law fashion as we increase model size, dataset size, and compute (Issue #170 – Data and Parameter Scaling Laws for Neural Machine Translation). In other words, bigger is better, and not just linearly better but following certain scaling exponents. These scaling laws mean that if you want to reduce the loss (or improve perplexity) by a certain amount, you can estimate how many more parameters or training tokens you’d need. This guided the creation of models like GPT-3 – simply by following the trend, GPT-3 at 175 billion parameters was expected to perform much better than its predecessors. However, scaling laws also showed diminishing returns and the need to keep data size and model size in sync to not under-utilize either. In 2022, the Chinchilla paper (by DeepMind) refined this: they argued many models were undertrained relative to their size, and that for a given compute budget, a smaller model trained on more data actually performs better than a huge model on limited data. The Chinchilla rule-of-thumb was to scale data and model such that $N_{\text{params}} \approx N_{\text{training\ tokens}}$ (with some constant). For an LLM developer, these insights are gold – they inform how you allocate resources. If you have the ability to train, say, a 60B model, you better also have on the order of 60B tokens of high-quality data to feed it. Otherwise, you’d be better off with a 30B model on 60B tokens, etc. The high-level point: LLMs get better with scale, and understanding the empirical scaling laws helps in planning a training run or anticipating model improvements (Issue #170 – Data and Parameter Scaling Laws for Neural Machine Translation). It also means that many behaviors emerge only at a certain scale (for instance, GPT-3’s few-shot learning capability was much weaker in smaller models). As a beginner, you don’t need to derive these laws, but it’s useful to read the plots from Kaplan’s paper (Issue #170 – Data and Parameter Scaling Laws for Neural Machine Translation) – they show remarkably smooth trends, suggesting we’re not near fundamental limits yet (though some believe we’ll hit limits as we exhaust the “low-hanging textual knowledge” available online).

Reinforcement Learning from Human Feedback (RLHF): Large language models like GPT-3, when trained purely on text, can generate fluent text but may not always align with what users want – they might produce incorrect facts, or inappropriate content, or simply ramble. RLHF is a technique to fine-tune models using human feedback as a signal of quality, thereby aligning the model’s behavior with human preferences. The process usually goes like this (Reinforcement Learning From Human Feedback (RLHF) For LLMs):

  1. Collect human feedback data: Typically, you have the model generate several responses to a variety of prompts. Human annotators then rank these responses from best to worst, or at least indicate which of two responses is better.
  2. Train a Reward Model: Using this human feedback (the rankings), train a separate model (or the same model in a different head) to predict a reward score for a given model output. Essentially, the reward model learns to imitate the human preferences – it should give higher scores to outputs humans preferred.
  3. Fine-tune the original model with RL (e.g., PPO): Now, treat the language model as an agent that, given a prompt (state), produces a response (action). Use a reinforcement learning algorithm (Proximal Policy Optimization is commonly used) to adjust the model’s parameters such that it maximizes the reward model’s score for its outputs (Reinforcement Learning From Human Feedback (RLHF) For LLMs). During this process, to avoid the model going off-distribution too much (since the reward model is an imperfect proxy and could be exploited), techniques like KL-divergence penalties (keeping the fine-tuned model close to the original pre-trained model) are used.
  4. The end result is a model that tries to produce outputs that humans would rate highly.

ChatGPT’s training is a prime example of RLHF in action: Starting from a GPT-3.5 base, they performed supervised fine-tuning on some prompts & ideal answers, and then did RLHF where humans preferred more helpful, correct, and harmless answers (Reinforcement Learning From Human Feedback (RLHF) For LLMs). The RLHF process significantly improved the helpfulness and safety of the model. As a learner, understanding RLHF is important because it addresses a key limitation of LLMs: the training objective (predict next token) is not directly our objective (give helpful, correct answers). RLHF is a method to bridge that gap using human intuition. It brings in reinforcement learning concepts like reward and policy optimization into NLP. If you build your own LLM for, say, a chatbot, you might consider collecting user feedback and applying RLHF to fine-tune it.

Building and Using LLMs: Practical Considerations

Putting it All Together – Training an LLM: Suppose you wanted to build a GPT-like model from scratch as Karpathy demonstrates in his “minGPT” project. You would:

  • Prepare a large corpus of text (could be scraped or from sources like Wikipedia, books, etc.). Clean and tokenize the text.
  • Set up a Transformer decoder model with a certain number of layers, attention heads, and model dimension. Initialize parameters (often randomly or with small random values).
  • Train the model to minimize cross-entropy loss on predicting the next token for each position in your text sequences. Use an optimizer like Adam with a learning rate schedule (Karpathy often emphasizes the importance of tuning these hyperparameters carefully (Neural Networks: Zero To Hero)).
  • Over weeks of training on GPUs/TPUs, monitor the training and validation loss. Possibly adjust things if it plateaus (e.g., learning rate adjustments).
  • Evaluate the trained model’s perplexity, and also test it qualitatively by prompting it to generate text.
  • If the base model is pre-trained well, you might then fine-tune it on a specific domain or with RLHF for alignment.

While doing this at the GPT-3 scale is out of reach for most individuals (it requires huge compute), doing a smaller scale version is educational. Karpathy, for instance, trains character-level language models on Shakespeare or GitHub code in his videos (Neural Networks: Zero To Hero) (Neural Networks: Zero To Hero), which is something you can replicate on a single GPU.

Prompt Engineering: Once you have a trained LLM, how you use it to get desired behavior is an art in itself. Prompt engineering is the process of designing the input text (prompt) to guide the model to produce the best output for your task (Prompt Engineering for AI Guide | Google Cloud). Since the model is essentially a next-word predictor, the only way we “program” it is through the prompt. For example:

  • If you want a model to act as a translator, you might prompt: “Translate the following English sentence to French: ‘Where is the library?’ ->”. The prompt includes an explicit instruction and perhaps an arrow or something to indicate where the answer should go.
  • If you want a summary, you might say: “Text: \n\nSummary in one sentence:”.
  • Few-shot prompting: Provide a few examples in the prompt. E.g., “Q: 2+2\nA: 4\nQ: 3+5\nA: 8\nQ: 7+10\nA:” – this shows the model examples of a task (here, addition) and primes it to continue in that pattern for the next query.

A well-crafted prompt can drastically improve the output quality of an LLM. Prompt engineering has become somewhat of a skill, with the best prompts often discovered through trial and error. It involves choosing the right wording, providing context, maybe constraints (“Answer in JSON format.”), or role-playing instructions (“You are an expert travel guide.”) – anything that helps the model understand what kind of output is desired (The Art of Prompt Engineering: Crafting Conversations with AI) (Prompt Engineering Explained: Crafting Better AI Interactions). Given that LLMs like GPT-3 have no explicit built-in notion of tasks, the prompt has to contain all the information about what task to perform.

For beginners, a fun exercise is to take a smaller pre-trained model (like GPT-2) and try different phrasing of prompts to see how the outputs differ. You’ll find that even small changes can influence the style or accuracy of responses. There’s also a growing body of knowledge (blog posts, papers) on prompt engineering techniques – for instance, chain-of-thought prompting, where you prompt the model to produce its reasoning step by step (like “Let’s think step by step”) which often leads to better logical results on math or reasoning problems by encouraging the model to articulate an intermediate reasoning process.

Prompt engineering is especially important when you can’t fine-tune the model (like when using an API) or when serving different queries dynamically. It’s also become somewhat the default way non-technical end-users interact with LLMs (through carefully phrased queries). So, mastering it is quite useful. Google Cloud’s guide says it well: prompt engineering is the art of designing prompts to guide AI models (especially LLMs) toward generating the desired responses (Prompt Engineering for AI Guide | Google Cloud).

Ethics and Bias: Although not explicitly in the outline, any comprehensive study plan for LLMs should include understanding the ethical implications. LLMs can emit biases present in training data or produce harmful content. Techniques like RLHF are partly aimed at mitigating this, but one also needs awareness (e.g., models might stereotype or might output false information in a confident tone). Andrej Karpathy has noted the importance of responsible AI usage. When building models, consider dataset curation (filter out highly toxic content, etc.) and testing the model for biases or weaknesses. As a beginner, an exercise could be to see how a model completes sentences like “The doctor said that…” vs “The nurse said that…” to observe if it has gender biases, for example.

Keep Reading and Experimenting: This field moves fast. While this report lays out the foundations, staying up-to-date via blogs (like Karpathy’s own blog or OpenAI’s research papers), courses (e.g., DeepLearning.AI’s Transformer specialization), and hands-on projects is key. As Karpathy famously did, don’t be afraid to read the classics (papers like “Attention is All You Need”) but also dive into code. By reimplementing things from scratch (a Karpathy hallmark), you solidify your understanding. For instance, writing a simple backpropagation by hand (as in Karpathy’s nano-grad) (Neural Networks: Zero To Hero) or coding a minified transformer can be incredibly educational.

IV. Thinking and Learning Strategies

Becoming proficient in the above areas is a significant undertaking. Here are some learning strategies and mindsets – many inspired by Karpathy’s own approach to learning – to help you on this journey:

First-Principles Thinking

Adopt a first-principles thinking approach to problems – break them down to fundamental truths and reason from the ground up (First Principles: The Building Blocks of True Knowledge). Rather than memorizing formulas or blindly following frameworks, try to understand why things are the way they are. For example, when learning backpropagation, derive the gradients yourself for a simple 2-layer network; when learning about transformers, step through a toy example of attention by hand. First-principles thinking will help you troubleshoot and innovate. Karpathy exemplifies this by building neural nets from scratch in Python, demystifying each component. By reducing complex problems to basic elements, you can reassemble solutions in your own way (First Principles: The Building Blocks of True Knowledge). This approach also helps in interviews and research, where being able to derive or reason things out is more valuable than rote knowledge.

Metacognition and Curiosity

Metacognition – or “thinking about your thinking” – means being aware of how well you understand something and regulating your learning strategies accordingly (Metacognition – Wikipedia). As you study, constantly ask yourself: Do I really get this? If you can’t explain it simply (even just to yourself), you probably need to revisit it. Karpathy often encourages understanding concepts to an intuitive level. Cultivate a habit of self-explanation (even if just via pen and paper). If a concept is fuzzy, identify the gap (“I understand how to compute an eigenvalue, but I’m not sure what it means geometrically”) and address it. Metacognition also involves planning your learning (“I’ll spend two weeks on linear algebra basics before moving on”) and monitoring progress (“I solved 10 practice problems on gradients correctly, so I think I’ve got it”) (Metacognition in the Classroom: Benefits & Strategies) (Metacognition – Wikipedia). Being reflective about your learning process will make you more efficient.

Additionally, stay curious. Read broadly about successes and failures in AI. If something intrigues or confuses you, chase down the answer. This might mean reading research papers or implementing a mini experiment. Curiosity-driven learning often leads to deeper understanding and more enjoyment.

Active Recall and Spaced Repetition

When studying theory or reviewing new material, use active recall – actively quiz yourself, rather than passively re-reading notes. For instance, after reading about backprop, close the book and try to write out the backprop equations from memory, or explain it aloud. Active recall strengthens memory by forcing your brain to retrieve information, which is much more effective than passively seeing it again (What Is Active Recall? – OnlineMedEd). Tools like flashcards (Anki) are great for this, especially for remembering definitions or equations (e.g., “What’s the formula for cross-entropy loss?” or “What does KL-divergence measure?”).

Combine this with spaced repetition – reviewing material at increasing intervals (the spacing effect) to ensure it moves to long-term memory (Spaced repetition (article) | Learn to Learn | Khan Academy). For example, review a concept one day after learning, then 3 days later, then a week, then a month. Each time, try to actively recall it. This is especially useful for math basics or terminologies that you might not use every day but are important. Over months, spaced repetition can greatly solidify your foundation knowledge so that it’s readily available when tackling advanced topics.

Project-Based Learning

Theory sinks in best when applied. Engage in project-based learning: pick small projects that interest you and force you to apply what you’ve learned. For instance, after learning about CNNs, you might implement your own digit recognizer for MNIST or a convolutional autoencoder. After learning transformers, you could train a small transformer to generate song lyrics. Projects provide context and motivation – they turn abstract concepts into concrete results. They also teach practical skills (loading data, debugging, tuning hyperparameters) that pure theory study might not. As the saying goes, you learn by doing. Karpathy’s career is punctuated with projects – from his early convnetJS demo (a CNN in JavaScript) to later real-world systems at Tesla. Start with manageable scopes so you can finish and feel a sense of accomplishment, then gradually increase ambition. The experience of completing a project will reinforce the knowledge far more than reading about it alone.

Project-based learning also gives you a portfolio to showcase, and it mirrors how real-world learning works (you encounter a problem, you figure out what you need to learn to solve it). It’s okay if you don’t know everything at the start of a project – that’s how you identify what you need to learn next.

Read Widely and Learn from Others

Don’t silo yourself to one resource or perspective. Reading widely exposes you to different explanations and viewpoints, which can deepen and broaden your understanding (Why Reading Widely is More Important Than Ever Before | by Consistent Climb | Medium). For example, you might read one textbook’s chapter on optimization, watch a YouTube lecture by Andrew Ng on it, and read a blog post by Karpathy – each might give a different intuition or highlight different aspects. Similarly, read papers and articles beyond just the technical how-tos: for instance, the original “Attention is All You Need” paper, or OpenAI’s blogs on GPT-3, or DeepMind’s papers on AlphaGo. These not only teach concepts but also the thought process of top researchers. Reading widely (from academic papers to online forums like Reddit or Hacker News discussions) also keeps you aware of emerging trends and common pitfalls, and challenges your assumptions (Why Reading Widely is More Important Than Ever Before – Medium). It broadens your perspective – as Mortimer Adler said, “You must be able to read widely if you are to be intellectually complete…” (Why Reading Widely is More Important Than Ever Before | by Consistent Climb | Medium). By seeing the same concept explained in multiple ways, or seeing different applications of a tool, you form a more well-rounded understanding.

Also, engage with the community: follow AI researchers and educators on Twitter (Karpathy’s Twitter is famously insightful and humorous about AI), join communities (the Fast.ai forums, Reddit’s r/MachineLearning or r/LanguageTechnology, etc.), attend online workshops or meetups if you can. Discussing with others or even just lurking and reading discussions can expose you to questions you didn’t think to ask and tips you wouldn’t find in a book.

Continual Learning and Big-Picture Thinking

Always remember the big picture: it’s easy to get lost in details of hyperparameters or specific model architectures. Regularly zoom out and remind yourself of the ultimate goals – e.g., building an AI system that can understand and generate language as well as a human (or better). This helps maintain motivation and context. Think about how the pieces connect: math enables ML algorithms; ML algorithms enable models like transformers; transformers enable applications like chatbots or translation systems.

Finally, accept that this is a continual learning field. What’s state-of-the-art now might be outdated in a couple of years. The mindset to cultivate is one of lifelong learning – the tools you acquire now (first-principles reasoning, mathematical rigor, coding skills, etc.) will allow you to learn new things down the line. Karpathy himself, after a stint in industry, came back to focusing on fundamental research/education, showing that one often circles back to learning new fundamentals for new problems.

In summary, approach learning AI/LLMs with the same curiosity and rigor that you approach building them:

  • Break problems down and understand them from scratch (first-principles).
  • Reflect on your own understanding (metacognition) and fill gaps actively.
  • Use proven study techniques like active recall and spaced repetition for theoretical parts.
  • Do projects to apply knowledge and make it stick.
  • Read and engage widely to broaden your horizons (as one blog put it, when we read widely, we expose ourselves to a range of perspectives and ideas, expanding our understanding and challenging our assumptions (Why Reading Widely is More Important Than Ever Before | by Consistent Climb | Medium)).
  • And keep tinkering – the field of AI is as much empirical as theoretical, and a lot of insight comes from experimentation.

Conclusion and Next Steps: You now have a roadmap of foundational topics – from linear algebra and calculus through classical ML to transformers and modern LLM techniques – and strategies to master them. A sensible path might be:

  1. Strengthen your math (perhaps via the suggested resources like the Essence of Linear Algebra videos (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.) or Khan Academy courses (GitHub – mlabonne/llm-course: Course to get into Large Language Models (LLMs) with roadmaps and Colab notebooks.)).
  2. Take an introductory ML course (Andrew Ng’s Machine Learning or Deep Learning Specialization are great starts) to get familiar with regression, SVMs, etc.
  3. Work through a deep learning course (like CS231n for CNNs or the fast.ai course, or Karpathy’s Neural Networks: Zero to Hero series (Neural Networks: Zero To Hero)) to understand training neural nets.
  4. Implement a project like training a small character-level language model (Karpathy’s makemore tutorial in his videos is a perfect guidance (Neural Networks: Zero To Hero)).
  5. Study the transformer architecture (perhaps start with the illustrated guide by Jay Alammar or the Dive into Deep Learning chapter on attention) and try implementing a mini-transformer or use Hugging Face Transformers library to fine-tune a model on some data.
  6. Learn about and perhaps implement a simple version of RLHF (maybe on a smaller problem, like training an agent with human feedback in a simulator, or just conceptually understand it from OpenAI’s summaries).
  7. Throughout, keep revisiting math (you’ll appreciate linear algebra more when you see it in backprop, and probability more when you see it in cross-entropy and attention distributions).

Remember to pace yourself – this is a marathon, not a sprint. But with consistent effort and the strategies above, you’ll steadily transform from a motivated beginner to an expert capable of building and understanding large language models. Good luck on your learning journey, and as Andrej Karpathy often encourages: stay inspired and keep coding!

Sources and Further Reading:

By following this roadmap and utilizing these resources, you’ll be well on your way to mastering the foundations needed to build large language models. Keep the learning active and enjoyable, and soon you’ll be creating impressive NLP projects of your own. Happy learning!

Glossary of Technical Terms

Active Recall

A study technique where you actively retrieve information from memory rather than passively reviewing it. Examples include flashcards or quizzing yourself without looking at notes.

Adam Optimizer

A popular optimization algorithm for training neural networks, combining ideas from momentum and adaptive learning rates. It maintains moving averages of both the first (mean) and second (variance) moments of gradients to adjust each parameter’s learning rate.

Algebra

A branch of mathematics dealing with symbols and the rules for manipulating those symbols. In machine learning, algebraic manipulations (like solving for unknowns in equations) appear often in regression, transformations, and matrix factorization.

Autograd (Automatic Differentiation)

A mechanism provided by deep learning frameworks (like PyTorch) that automatically computes gradients of tensors with respect to given operations. This is how neural networks efficiently implement backpropagation without the programmer manually coding the gradient equations.

Backpropagation

The fundamental algorithm used to train neural networks by calculating gradients of the loss function with respect to the network’s parameters. It works by applying the chain rule of calculus layer by layer from the output back to the input.

Batch Normalization

A technique that normalizes activations in intermediate layers of a neural network to stabilize and speed up training. It can also have a slight regularizing effect, helping the model generalize better.

BERT (Bidirectional Encoder Representations from Transformers)

A Transformer-based model that uses an encoder-only architecture. Trained primarily through masked language modeling (predicting randomly masked words), it captures bidirectional context, making it powerful for understanding-oriented NLP tasks.

Big-O Notation

A notation describing the time complexity (or space complexity) of an algorithm in terms of input size. For instance, O(n) means the run time grows roughly linearly with input size n; O(n²) means it grows roughly with the square of n.

Bidirectional Context

In language models, this means the model can “look” at both previous and subsequent words in a sentence to understand the meaning, rather than only the past. BERT’s encoder is fully bidirectional.

Chain-of-Thought Prompting

A prompting strategy in LLMs where the model is encouraged to generate intermediate reasoning steps (like a written chain of thought) before giving a final answer. This can improve the accuracy of tasks requiring multi-step reasoning.

Chinchilla

A guideline/model suggesting that for a given compute budget, one should balance model size and the number of training tokens. It refines earlier scaling laws and shows that many prior large models were undertrained relative to their size.

Classification

A core machine learning task that assigns an input (image, text, etc.) to one of several discrete classes or categories (e.g., spam vs. not spam).

CNN (Convolutional Neural Network)

A neural network architecture specialized for grid-like data such as images. It uses convolutional layers that apply learnable filters, detecting local patterns (edges, shapes) as they slide over the input.

Computer Architecture

The design of computer systems (CPUs, GPUs, memory hierarchies, etc.). Understanding architecture helps optimize software (like machine learning training) for speed, parallelism, and efficient resource usage.

Cross-Entropy Loss

A loss function commonly used in classification and language modeling. It measures the difference between the predicted probability distribution and the true distribution. Minimizing cross-entropy is equivalent to maximizing the likelihood of the correct label or token.

Data Structures

Ways to organize and store data in computers (arrays, linked lists, trees, hash tables, etc.). They affect how efficiently we can access and modify data, which is crucial in large-scale ML/NLP pipelines.

Decision Tree

A tree-structured model that splits data based on feature thresholds, leading to leaf nodes that make predictions. It is intuitive but can overfit if not pruned or if grown too deeply.

Derivative

In calculus, the derivative of a function indicates its instantaneous rate of change. In machine learning, derivatives form the basis of gradient-based optimization (backpropagation).

Dimensionality Reduction

Techniques (like PCA) that reduce high-dimensional data to fewer dimensions, often to visualize or remove noise/redundancies. It uses concepts like eigenvalues/eigenvectors or more advanced manifold learning methods.

Dropout

A regularization technique in neural networks where random neurons are “dropped” (set to zero) during training. It prevents over-reliance on specific neurons and thus helps reduce overfitting.

Encoder-Decoder Architecture

A neural network design, often used for tasks like machine translation, where an encoder processes the input into a context representation, and a decoder generates the output sequence from that context.

Entropy (Information Theory)

A measure of uncertainty or “surprise” in a random variable. Higher entropy means outcomes are more unpredictable. In ML, cross-entropy and perplexity are related metrics for measuring predictive uncertainty.

Evaluation Metrics

Criteria or quantitative measures used to assess a model’s performance, e.g., accuracy, F1-score for classification, perplexity for language models, or mean squared error for regression.

Feed-Forward Neural Network (MLP)

A basic neural network architecture where each layer is fully connected to the next and data flows in one direction. Typically used for simpler tasks or as a baseline model.

Few-Shot Learning

The phenomenon where a large language model can generalize to a new task given only a few examples in the prompt without additional parameter updates.

First-Principles Thinking

Approaching complex problems by reducing them to their most fundamental truths, then reasoning upward from there, rather than relying on analogy or memorization.

Gated Recurrent Unit (GRU)

A variant of RNN that uses gating mechanisms (reset and update gates) to better capture long-term dependencies than vanilla RNNs, but with fewer parameters than LSTMs.

Generative Pre-trained Transformer (GPT)

A decoder-only Transformer model that predicts the next token in a sequence (autoregressive). GPT models (e.g., GPT-2, GPT-3) learn language patterns from massive text data and are strong in text generation tasks.

Gradient

A vector of partial derivatives indicating how a function (e.g., a loss function) changes with respect to each parameter. Used in gradient-based optimization to update parameters in the negative direction of the gradient (reducing loss).

Gradient Descent

A fundamental optimization algorithm that repeatedly moves parameters in the opposite direction of the gradient to minimize the loss function. Stochastic Gradient Descent (SGD) uses batches of data for efficiency.

Hugging Face Transformers

A popular Python library that provides pre-trained NLP models (BERT, GPT-2, T5, etc.), tokenizer implementations, and training utilities, making it easier to experiment and fine-tune large language models.

Hyperparameters

Configuration variables in a model or learning process that are not directly learned from data (e.g., learning rate, number of layers, batch size). Tuning hyperparameters can drastically affect performance.

Information Gain

In decision trees, the reduction in entropy (uncertainty) about the target variable by splitting the data on a certain feature. The feature yielding the highest information gain is chosen for the tree split.

L1/L2 Regularization

Techniques adding a penalty term to the loss function based on the magnitude of the model’s weights. L1 sums the absolute values (promoting sparsity), and L2 sums the squares (promoting small, but not necessarily zero, weights).

Language Model

A model that learns the probability distribution over sequences of tokens (words, subwords, etc.). It can be used to predict the next token, generate text, or provide embeddings for downstream tasks.

Learning Rate

A hyperparameter in gradient-based optimization determining the size of each parameter update step. Too high can cause divergence, too low can cause slow convergence.

Linear Algebra

A branch of mathematics dealing with vectors, matrices, and linear transformations. Underpins almost all aspects of machine learning, particularly the representation of data and model parameters.

Linear Regression

A foundational method for predicting a continuous output by fitting a linear relationship (y = w·x + b). Though simple, it introduces core concepts like gradient descent and loss minimization.

Loss Function

A function measuring how far off model predictions are from the ground truth. Lowering loss improves model accuracy/performance. Examples include mean squared error (MSE) for regression and cross-entropy for classification.

LSTM (Long Short-Term Memory)

A variant of RNN designed to better capture long-range dependencies by using gates (input, output, forget) and a cell state, mitigating the vanishing/exploding gradient issues of vanilla RNNs.

Matrix

A rectangular array of numbers (arranged in rows and columns). Used to represent data, transformations (like neural network weights), and more. Key operations include matrix multiplication, inversion, decomposition, and eigenanalysis.

Metacognition

Awareness of and reflection on one’s own thinking and learning processes. Involves monitoring comprehension, identifying knowledge gaps, and regulating study strategies accordingly.

Mindmap

A diagram or visualization showing the relationships between key topics or concepts, often branching out hierarchically from a central theme.

Momentum

An enhancement to gradient descent that accumulates an exponentially decaying moving average of past gradients, giving updates “inertia” and smoothing out oscillations.

MSE (Mean Squared Error)

A common loss function for regression problems. It calculates the average of the squares of the differences (errors) between predicted and actual values.

MLP (Multi-Layer Perceptron)

Another name for a feed-forward neural network with at least one hidden layer containing nonlinear activation functions.

NLP (Natural Language Processing)

A field focused on enabling computers to understand, interpret, and generate human language. Spans tasks like text classification, translation, question answering, and more.

NumPy

A Python library providing support for large, multi-dimensional arrays and matrices, along with a collection of high-level mathematical functions to operate on them efficiently. Essential for scientific computing and ML tasks.

Optimizer

Algorithm or method used to adjust neural network parameters to minimize the loss function (e.g., SGD, Adam, RMSProp). Plays a central role in how fast and effectively a model learns.

Overfitting

When a model memorizes training data rather than learning a generalizable pattern. This usually leads to high training accuracy but poor test or validation accuracy.

PCA (Principal Component Analysis)

A dimensionality reduction method using linear algebra (eigenvalue decomposition of the covariance matrix) to find directions (principal components) capturing maximum variance in data.

Perplexity

An evaluation metric for language models, defined as the exponentiated average negative log-likelihood of the model’s predictions. Lower perplexity indicates the model is less “surprised” by the actual text.

Position Embeddings

Vectors added to token embeddings in a Transformer to encode the token’s position in a sequence. Necessary because self-attention alone does not inherently capture sequence order.

Probability & Statistics

The mathematical foundation for dealing with uncertainty, random variables, distributions, hypothesis testing, and inference. Critical in understanding data sampling, model evaluation, and Bayesian methods.

Prompt Engineering

The process of crafting specific input text prompts to guide large language models toward producing desired outputs. May include formatting instructions, context, or examples.

PyTorch

A deep learning framework that provides automatic differentiation (autograd), flexible GPU acceleration, and a Pythonic interface. Widely used for research and production ML systems.

Q, K, V (Query, Key, Value)

Vectors used in the attention mechanism of Transformers. Each token is mapped to a query, a key, and a value. The attention weights come from query-key dot-product alignment, and the actual output is a weighted sum of values.

Recurrent Neural Network (RNN)

A neural network type that processes sequences one element at a time while maintaining a hidden state, enabling it to capture temporal or sequential dependencies.

Regularization

Techniques to reduce overfitting and improve generalization of models. Can be added to the loss function (L1/L2), or introduced by structural methods (dropout, data augmentation, batch norm, etc.).

Reinforcement Learning from Human Feedback (RLHF)

A technique to align model behavior with human preferences by using human-labeled data to train a reward model, then fine-tuning the original model via a reinforcement learning algorithm to maximize that reward.

Residual Connections

Connections in deep neural networks that add the input of a layer directly to its output, helping gradients flow better and mitigating vanishing gradient issues. Common in modern architectures like ResNet and Transformers.

Scaling Laws

Empirical relationships indicating how model performance improves as model size, dataset size, and compute increase. They guide choosing model capacity and data quantity to maximize performance under a given compute budget.

Self-Attention

The key component in Transformers: each token attends to every other token in the sequence. By computing similarities (Q·K) and applying weights to values (V), the model captures relationships without relying on recurrence or convolution.

SGD (Stochastic Gradient Descent)

A variant of gradient descent that uses mini-batches of data to estimate the gradient for each parameter update. Faster and more memory-efficient than batch gradient descent on large datasets.

SVM (Support Vector Machine)

A classical machine learning algorithm that finds the hyperplane maximizing the margin between different classes (for classification). It also has variants for regression (SVR) and uses kernel methods for non-linear decision boundaries.

T5 (Text-to-Text Transfer Transformer)

A Transformer model using an encoder-decoder architecture, trained in a text-to-text paradigm on multiple tasks. It can be prompted to handle translation, summarization, and other NLP tasks by treating them all as “generate text” scenarios.

Tokenization

The process of splitting text into tokens (subwords, words, or characters) for input into an NLP model. Modern methods (e.g., Byte-Pair Encoding) handle large vocabularies efficiently and handle unknown words by splitting them further.

Transformer

A neural network architecture that relies on self-attention (and possibly cross-attention) rather than recurrence or convolution. Forms the basis for modern large language models like GPT and BERT.

Vector

An array of numbers arranged in a single row or column. In ML, vectors often represent data points, weights, or embeddings. Operations include addition, scalar multiplication, and dot products.

Word Embeddings

Continuous vector representations of tokens (words/subwords) learned by a model. Similar words or tokens tend to have similar embeddings in semantic space.

Share via
Copy link