Fact-checked by Grok 2 weeks ago

Compiler

A compiler is a computer program that translates source code written in a high-level programming language, such as C or Java, into a lower-level target language, typically machine code or bytecode, to produce an executable program that a computer can run directly. This translation enables developers to write abstract, human-readable instructions without needing to manage low-level hardware details, fundamentally separating programming logic from machine-specific operations. The architecture of a compiler is divided into a front-end and a back-end, with the front-end handling language-specific tasks and the back-end focusing on machine-specific output. Key phases in the front-end include , which breaks the source code into tokens like keywords and identifiers while ignoring whitespace and comments; syntax analysis, which parses tokens to build an representing the program's structure; and semantic analysis, which checks for meaning and context, such as type compatibility and scope resolution. In the back-end, intermediate code generation creates a platform-independent representation, followed by optimization to improve efficiency—such as eliminating redundant computations—and final to produce target machine instructions. The development of compilers traces back to the early , when the term "compiler" was coined by pioneering computer scientist Grace Murray Hopper to describe a system for automatically translating high-level instructions into . One of the first practical implementations was the compiler in the late 1950s, which demonstrated the feasibility of generating optimized from a problem-oriented source language, marking a shift from hand-written to automated translation. This innovation laid the groundwork for modern by enabling portable, efficient code across diverse hardware. Compilers play a pivotal role in by enhancing software , , and through advanced optimizations and detection, while also facilitating the creation of domain-specific languages and tools. Unlike interpreters, which execute code line-by-line at , compilers produce standalone executables that run faster and with lower resource overhead, making them indispensable for large-scale applications. Ongoing research continues to evolve compilers to support emerging paradigms like and AI-driven .

Overview

Definition and Purpose

A compiler is a computer program that translates source code written in a high-level programming language into a lower-level target language, such as machine code or bytecode, suitable for execution on a specific computer architecture. This translation process ensures that the resulting program maintains the intended functionality of the original source while adapting it to the hardware's instruction set. The primary purpose of a compiler is to bridge the gap between human-readable, abstract programming languages and the low-level instructions required by processors, thereby enabling developers to focus on algorithmic logic rather than machine-specific details. By abstracting complexities, compilers promote portability across different systems and improve overall software efficiency through optimizations that produce faster, more resource-efficient executables. This abstraction not only accelerates but also enhances and in large-scale applications. At a high level, a compiler's involves taking as input, performing an analysis phase to parse and validate the code's structure and semantics, followed by a phase to generate the target output, such as or an binary. For instance, the early developed by in 1952 represented one of the first such translation tools, functioning as a rudimentary compiler for the computer by linking subroutines into form. A practical example is compiling a simple C program, such as one printing "Hello, World!", using a tool like GCC with the -S flag to produce x86 assembly code; this generates instructions like mov and call that the assembler then converts to machine code for execution on an x86 processor.

Key Components

A compiler's internal structure is typically modular, comprising several core components that process source code through distinct stages to produce executable target code. The primary components include the lexical analyzer (also known as the scanner), parser, semantic analyzer, code generator, and symbol table manager. These elements work in sequence to analyze and synthesize code, ensuring syntactic and semantic correctness while optimizing for the target machine. The lexical analyzer, or , is the first component in the , responsible for reading the source code as a stream of characters and converting it into a sequence of meaningful tokens, such as keywords, identifiers, operators, and literals. It performs tasks like removing comments and whitespace, expanding macros, and handling language-specific features such as or indentation-based scoping (e.g., in ). By grouping characters into tokens, the simplifies subsequent and detects basic lexical errors, such as invalid characters. Following the scanner, the parser, or syntax analyzer, takes the token stream as input and constructs an (AST) or that represents the hierarchical structure of the program according to the language's . It verifies that the tokens conform to syntactic rules, resolving ambiguities through techniques like or LR parsing, and reports syntax errors if the structure is invalid. The resulting AST captures the program's logical organization, such as expressions, statements, and declarations, enabling higher-level analysis. The semantic analyzer then processes the to enforce static semantic rules beyond syntax, such as type checking, scope resolution, and declaration consistency. It verifies that operations are semantically valid—for instance, ensuring no type mismatches in expressions or undeclared variable usage—and annotates the with additional information, like inferred types or resolved references, often producing a decorated or semantic graph. This phase detects errors that syntax alone cannot catch, such as assigning a to an integer variable. The code generator operates toward the end of the process, translating the semantically verified (IR)—typically generated after semantic analysis—into target-specific or . It maps high-level constructs to low-level instructions, incorporating and addressing modes tailored to the target architecture, such as or , to produce efficient executable output. Throughout these stages, the manager maintains a dynamic that stores and retrieves information about program symbols, including variables, functions, types, and scopes. It tracks attributes like declaration locations, types, and , supporting lookups during , semantic , and to resolve references and prevent errors like redeclarations. The facilitates interconnections by providing shared access to symbol data across components, often implemented as hash tables or trees for efficient operations. These components interconnect in a linear data flow: the source code passes to the lexical analyzer, yielding tokens to the parser, which builds the for the semantic analyzer; the enriched then feeds the code generator, with the manager integrating data bidirectionally to support analysis and synthesis. This allows for independent development and reuse, as seen in compiler toolkits like Lex for generating lexical analyzers from specifications and for producing parsers from context-free grammars, which accelerate prototyping by automating routine tasks. A typical block diagram of these interactions can be visualized as follows:
[Source Code](/page/Source_code)
     |
     v
Lexical Analyzer ([Scanner](/page/Scanner))
     |
     v
[Tokens](/page/The_Tokens)
     |
     v
Parser
     |
     v
Abstract Syntax Tree (AST)
     |
     v
Semantic Analyzer
     |          ^
     v          | (Symbol lookups)
[Intermediate Representation](/page/Intermediate_representation) (IR) <--- [Symbol Table](/page/Symbol_table) Manager
     |
     v
Code Generator
     |
     v
Target Code
This flow highlights the progressive transformation from raw input to executable output, with the symbol table enabling cross-phase consistency.

Comparison with Other Systems

Interpreters

An interpreter is a program that executes directly, typically line by line or statement by statement, without first translating the entire code into a separate file. Unlike compilers, which produce for later execution, interpreters read, analyze, and run the code on-the-fly during . The execution model of an interpreter involves processing the source sequentially, often translating it into an intermediate form such as before immediate execution, sometimes within a environment that abstracts the underlying hardware. This approach allows for dynamic evaluation, where can be modified or extended while running, providing greater flexibility for interactive and scripting applications. Interpreters offer several advantages over compilers, including easier due to line-by-line execution that pinpoints errors immediately, with instant feedback without a separate compilation phase, and enhanced portability across platforms via the layer. For instance, Python's implementation exemplifies this model, where source code is compiled to and then interpreted by the , enabling seamless execution in interactive sessions or scripts. However, interpreters generally incur performance penalties, as the repeated and at each run eliminate opportunities for whole-program optimizations that compilers can apply, resulting in significantly slower execution speeds, with slowdown factors often ranging from 10x to thousands compared to compiled equivalents, depending on the workload. This overhead stems from the on-the-fly processing, which avoids upfront compilation but limits efficiency for compute-intensive tasks. Earlier examples include Short Code, proposed in 1949 by John Mauchly and implemented for computers like the BINAC and UNIVAC I, which used an interpreter to execute high-level instructions line by line.) A notable early interpreter was developed for the Lisp programming language starting in fall 1958 by John McCarthy and his team at MIT, initially as a simpler alternative to building a full compiler, marking a pivotal innovation in symbolic computation and AI research. Hybrid approaches, such as those combining interpretation with just-in-time compilation, bridge the gap by interpreting initially for quick startup while compiling hot code paths for improved performance, though these extend beyond pure interpretation.

Assemblers and Linkers

Assemblers are specialized programs that translate assembly language, a low-level symbolic representation of machine instructions, into relocatable machine code. Assembly language uses mnemonic instructions and symbolic addresses to represent binary operations, providing a one-to-one correspondence with the target processor's instruction set architecture (ISA). This translation process involves resolving symbols, such as labels for memory locations, into actual addresses while generating object files that include the machine code and relocation information for later use. Unlike compilers, which process high-level source code through syntactic and semantic analysis, assemblers operate directly on this intermediate, ISA-specific input without performing complex optimizations or type checking, focusing instead on straightforward resolution and mapping. Typically, assemblers perform this in one or two passes: the first pass builds a and calculates addresses, while the second generates the by substituting machine for mnemonics. Linkers, also known as link editors, take the output from compilers or assemblers—object files containing relocatable and unresolved external references—and combine them into a single suitable for execution on the target system. Their primary functions include symbol resolution, where undefined references (e.g., calls to functions in other modules or libraries) are matched to their definitions; relocation, which adjusts addresses to fit the final memory layout; and , incorporating necessary routines from static or shared libraries. Linkers support two main types of linking: static and dynamic. In static linking, all required library code is copied into the executable at link time, resulting in a self-contained binary that does not depend on external files at runtime, though this increases file size. Dynamic linking, by contrast, defers resolution until load or run time, referencing shared libraries (e.g., .so files on Unix-like systems) to reduce redundancy and enable updates to libraries without recompiling the program. A prominent example is the GNU linker (ld), part of the GNU Binutils suite, which supports multiple object formats like ELF and uses a command language for customizing output sections and memory placement. In the typical compilation pipeline, a compiler generates from high-level source, which may already include assembly-like intermediates; if source is written in , an assembler explicitly produces the . The linker then processes these s, resolving dependencies and producing the final , ensuring the program can run independently or with dynamic libraries. Early assemblers emerged in the to simplify programming of the first electronic computers, with developing the initial , known as "Contracted Notation" or autocode, in 1947 for the Automatic Relay Computer (ARC) at Birkbeck College, . This innovation replaced tedious binary coding with symbolic instructions, paving the way for more efficient low-level programming before the widespread adoption of high-level languages.

Historical Development

Early Innovations (1950s–1970s)

The development of compilers began in the early 1950s amid significant hardware constraints, including limited memory and processing power on machines like the UNIVAC I, which compelled innovators to create tools for automating code translation from high-level descriptions to machine instructions. Grace Hopper led the creation of the A-0 system in 1952 at Remington Rand, recognized as the first compiler precursor, which automatically linked and ordered subroutines based on specifications read from punched cards, marking an initial step toward automatic programming. Concurrently, A.E. Glennie developed Autocode in 1952 for the Manchester Mark 1 computer, an early high-level programming system that translated mathematical expressions into machine code, addressing the tedium of manual assembly coding. By 1954, R.A. Brooker at the University of Manchester advanced this with the Mark 1 Autocode, introducing the first compiled subroutines that generated efficient machine code for reusable code blocks, overcoming manual optimization challenges on resource-scarce hardware. A pivotal milestone arrived in 1957 with the release of the compiler by and his team at for the , the first widely used high-level language compiler that translated scientific formulas into optimized code, dramatically reducing programming time from weeks to hours despite the era's 36-bit word size and 4K-word memory limits. This compiler incorporated early optimization techniques, such as and , to produce code rivaling hand-written in efficiency, thus proving high-level languages viable on limited . Key figures like , , and Brooker navigated these constraints by emphasizing syntax analysis and , laying groundwork for structured translation methods. In the 1960s, , formalized in 1960 by an international committee including , profoundly influenced by introducing block structures, recursive procedures, and lexical scoping, which compilers could enforce to promote readable, modular code free from unstructured jumps. This language spurred advancements in compiler design, including syntax-directed —a technique where drives semantic actions for —exemplified in early implementations that used recursive descent parsers to handle nested blocks efficiently. 's FORTRAN IV, released in 1962 for systems like the IBM 7090, built on prior versions with enhanced optimizations such as index register usage and , enabling better performance on evolving hardware while standardizing features like explicit type declarations to ease portability. The 1970s saw further innovation with developing the language in 1972 at , designed as a portable tool that combined high-level abstractions with low-level access, initially compiled for the PDP-11 to rewrite Unix in by 1973. Ritchie's collaboration with integrated into the emerging Unix , where the compiler formed a core component alongside assemblers and linkers, facilitating modular on minicomputers. To address machine dependency, Stephen C. Johnson's (PCC) in the mid-1970s employed an intermediate language representation, allowing front-end parsing to target multiple back-ends and enabling Unix porting across architectures like the VAX, thus overcoming hardware fragmentation through retargetable design. These efforts by Ritchie, Thompson, and Johnson resolved persistent challenges like manual optimization and non-portability, establishing compilers as essential for scalable software ecosystems.

Modern Advancements (1980s–Present)

The 1980s and 1990s saw the rise of open-source compilers that emphasized portability and modularity, addressing the limitations of proprietary systems tied to specific . The GNU Compiler Collection (), launched by in 1987 as part of the GNU Project, became a cornerstone for development by supporting multiple languages and architectures, enabling developers to build portable applications without . By the late 1990s, had evolved to support several programming languages, including C, C++, , , and Ada, and numerous processor architectures, fostering widespread adoption in systems and beyond. The Low Level Virtual Machine () project, started in 2000 by at the University of Illinois, introduced a modular compiler infrastructure that separated the front-end language parsing from back-end code generation, allowing reusable components across different languages and targets. This design facilitated easier experimentation with optimizations and supported the creation of compilers for new languages with minimal effort. In the 2000s, just-in-time () compilation gained prominence with the , released by in 1999, which dynamically compiled to native code at for improved performance in enterprise applications. Similarly, the .NET Framework's , introduced by in 2002, incorporated compilation to enable seamless execution of managed code across platforms, boosting the adoption of languages like C#. The decade also witnessed hybrid approaches for scripting languages, such as V8 for in 2008, blending with to balance startup speed and efficiency in web browsers. The brought compilers focused on safety and performance for modern paradigms. Mozilla's compiler (rustc), with development beginning in 2009 and the first stable release in 2015, pioneered memory-safe through its borrow checker, preventing common errors like dereferences without garbage collection, and has since been used in like the . Apple's compiler, announced in 2014, emphasized and high performance for development, compiling to native code while integrating with ecosystems to reduce boilerplate and errors. (Wasm) compilers emerged around 2015 as a portable binary format, with tools like (extended in this era) enabling C/C++ code to run efficiently in browsers via to Wasm modules. From the late to the , AI-driven optimizations transformed compiler design. LLVM's MLGO framework, integrated in 2021, uses to guide phase ordering and inlining decisions, achieving up to 1.3% geomean speedup on SPEC benchmarks compared to traditional heuristics. Key trends include modular architectures for easier extension, as seen in LLVM's influence on projects like and ; cross-compilation support for embedded and devices, with and enabling builds for and without host modifications; and the dominance of open-source compilers, which by 2023 powered over 90% of distributions and cloud infrastructure. In 2025, parallel compilation advancements in /LLVM, such as distributed task execution, improved build times for large projects, enhancing developer productivity. Quantum compilation also advanced, with IBM's framework incorporating circuit optimization passes in its 2025 release to map high-level quantum algorithms to noisy intermediate-scale quantum hardware more efficiently.

Compiler Architecture

Phases and Passes

Compiler processing is organized into phases and passes to systematically transform into executable . Phases represent distinct, sequential transformation steps that operate on the input, such as , which breaks the into tokens; syntax analysis, which verifies the structure against grammar rules; semantic analysis, which checks type compatibility and scope rules; and , which produces target code from an . These phases can be executed in a single traversal or repeated across multiple passes, depending on the compiler design. Passes, in contrast, refer to complete traversals of the source code or intermediate representations, where each pass may encompass one or more phases to build or refine the output progressively. One-pass compilers perform all necessary analysis and generation in a single sweep through the source code, enabling faster times but requiring languages with restricted features, such as no forward references to symbols or simple block structures to avoid deferred decisions. For instance, designed languages like Pascal to support single-pass for simplicity and efficiency in resource-constrained environments. In contrast, multi-pass compilers make multiple traversals, allowing phases to be separated and repeated for deeper analysis, such as multiple optimization iterations on intermediate code; this approach supports more complex languages and aggressive optimizations but increases time and usage due to repeated loading and processing. Early compilers like the I system from 1957 employed a multi-pass design with six passes to handle optimization despite limited hardware, marking a shift from manual assembly to automated code improvement. The trade-offs between one-pass and multi-pass designs balance compilation speed against code quality and language expressiveness. Single-pass approaches minimize overhead, making them suitable for interactive or systems where quick is essential, but they limit optimizations and error detection to what can be resolved on-the-fly. Multi-pass designs, while resource-intensive, enable comprehensive error checking and transformations, such as global , which are infeasible in a single traversal; however, they can lead to longer build times, especially in large projects. Over time, as hardware capabilities grew, compilers evolved toward multi-pass architectures to exploit opportunities for better performance, with modern systems like using dozens of passes for modular optimization. A simple example of multi-pass processing is the two-pass assembler, commonly used in early computing systems. In the first pass, the assembler scans the source to build a by assigning addresses to labels and resolving forward references without generating code. The second pass then traverses the code again, substituting symbol addresses and emitting machine instructions based on the table. This design illustrates the core benefit of passes: deferring resolution until sufficient information is gathered. Bootstrapping compilers, which compile their own , often rely on multi-pass structures to iteratively improve the compiler itself, starting from an initial version written in another language. The NELIAC compiler from 1958 pioneered this by using a bootstrap to generate subsequent versions on the 7090. In terms of structure, one-pass compilers follow a linear flow:
Source Code → Lexical → Syntax → Semantic → Code Gen → Machine Code
While multi-pass designs involve iterative or pipelined traversals, such as separate passes for (covered in the section) and (covered in the section), allowing reuse of .

Front End

The of a compiler performs the initial analysis of , verifying its adherence to the programming language's and semantics while generating a machine-independent . This phase is source-language dependent, focusing on transforming raw text into a structured form suitable for further processing. By isolating language-specific concerns, the enables reusability with multiple for different target architectures. The front end comprises three primary components: lexical analysis, syntax analysis, and semantic analysis. Lexical analysis scans the input stream of characters and groups them into meaningful units called tokens, such as keywords, identifiers, operators, and literals. This process relies on regular expressions to define token patterns, which are converted into deterministic finite automata (DFAs) for efficient recognition. For instance, the regular expression for an identifier, [a-zA-Z][a-zA-Z0-9]^*, matches a starting letter followed by zero or more alphanumeric characters, ensuring tokens like variable names are correctly isolated. Tools like Lex automate DFA construction from these specifications, optimizing scanning performance through state minimization. Syntax analysis builds upon the token stream to validate the program's structure against the language's , producing a that represents hierarchical relationships. Parsers operate either top-down, as in algorithms that predict expansions from the start symbol, or bottom-up, as in LR algorithms that reduce tokens to non-terminals via shift-reduce mechanisms. LR parsers, introduced by Knuth, handle a broader class of grammars deterministically in linear time, making them suitable for most programming languages. Parser generators such as , which produces LALR(1) parsers from grammar rules augmented with semantic actions, and its GNU successor , facilitate rapid development by automating table-driven parsing. Semantic analysis examines the for context-sensitive properties, ensuring the program's meaning is well-defined. Key tasks include type checking, which verifies operand compatibility in operations (e.g., disallowing of incompatible types like and ), and scope resolution, which binds identifiers to their declarations within nested scopes using mechanisms like symbol tables. Violations trigger errors, such as undeclared variables or type mismatches. This phase often integrates attribute grammars to propagate information across the tree. The front end culminates in an (AST), a condensed representation omitting extraneous details like parentheses, or a preliminary (IR) such as three-address code, where statements are simplified to forms like x = y + z. For the declaration int x = 5;, yields tokens for "int", "x", "=", and "5"; syntax analysis constructs nodes for a variable declaration with ; and semantic analysis confirms integer type assignment, yielding an AST with a declaration node parenting an identifier "x" and integer literal 5. To support robust compilation, the front end incorporates error handling and recovery strategies during lexical, syntactic, and semantic phases. Upon detecting issues like unexpected tokens, parsers may skip to synchronization points (e.g., semicolons) or insert/deleted symbols locally to resume , minimizing cascading errors and enabling multi-error reporting in a single pass.

Middle End

The middle end of a compiler, often referred to as the optimizer, bridges the front end and by transforming the (IR) into a more efficient form through machine-independent optimizations. These transformations focus on improving program performance, such as reducing computational redundancy or eliminating unnecessary operations, without relying on details of the target . Core techniques include , which evaluates expressions with known constant values at , and , which removes code that has no effect on the program's observable behavior, such as unreachable statements or computations whose results are never used. Central to middle-end optimizations are and s, which provide the analytical framework for identifying optimization opportunities. A (CFG) models the program's execution paths, with nodes representing basic blocks—sequences of instructions without branches—and edges indicating possible control transfers, enabling precise modeling of program structure for analysis. operates on this CFG to track variable states across the program; for example, reaching definitions analysis computes the set of variable definitions that may reach each program point, while live variables analysis determines which variables retain value beyond a given point. These analyses support optimizations by revealing redundancies, such as unused definitions that can be eliminated. The basic data-flow equations for reaching definitions illustrate this process formally. For a node n in the CFG: \text{RD}_{\text{in}}(n) = \bigcup_{p \in \text{pred}(n)} \text{RD}_{\text{out}}(p) \text{RD}_{\text{out}}(n) = \text{gen}(n) \cup \left( \text{RD}_{\text{in}}(n) - \text{kill}(n) \right) Here, \text{pred}(n) are the predecessors of n, \text{gen}(n) is the set of definitions generated within n, and \text{kill}(n) is the set of definitions invalidated by n. To solve, initialize all \text{RD}_{\text{in}} and \text{RD}_{\text{out}} sets to the (for a forward starting conservatively), then propagate information iteratively in across the CFG or using worklist algorithms until a fixed point is reached, where no further changes occur due to the monotonic nature of the union and set operations. This fixed-point computation ensures all possible reaching definitions are captured efficiently. Middle-end processing involves multiple optimization passes, often organized in loops to apply transformations iteratively as new opportunities arise from prior changes. For instance, (LICM) identifies computations within a loop that yield the same result on every iteration—such as loop-independent expressions—and relocates them to a pre-loop position, minimizing redundant execution while preserving program semantics. These passes rely on prior analyses like (nodes that precede all paths to a given ) to ensure safe movement. In practice, compilers execute sequences of such passes, potentially repeating them to handle interdependencies. Frameworks like LLVM's pass manager facilitate modular implementation of these optimizations, allowing developers to define and schedule sequences of and passes on the . The new pass manager in LLVM supports nested pipelines, enabling bottom-up optimization of call graphs and efficient reuse of results across passes, which enhances for large codebases. For example, the optimization of x = 2 + 3; to x = 5; via demonstrates a simple yet impactful , replacing with a direct assignment.

Back End

The back end of a compiler translates the optimized (IR) into target-specific , accounting for the hardware architecture's instruction set, registers, and memory model. This phase is highly dependent on the target platform, such as x86, , or , requiring architecture-specific rules to ensure efficient and correct execution. Unlike the machine-independent optimizations in prior phases, the back end focuses on low-level transformations to produce executable binaries or assembly code. Key components include instruction selection, register allocation, and code generation. Instruction selection maps operations in the IR, often represented as a directed acyclic graph (DAG), to concrete machine instructions using techniques like tree pattern matching. In this approach, the IR tree or DAG is covered by predefined patterns (tiles) corresponding to instruction alternatives, selecting the optimal match based on cost functions for latency or code size; for instance, the Bottom-Up Rewrite System (BURS) automates this by generating efficient parsers from a cost-augmented grammar. Register allocation assigns virtual registers from the IR to a limited set of physical machine registers, modeled as graph coloring on an interference graph where nodes represent live ranges and edges indicate overlapping lifetimes. If the graph is not colorable with the available registers, spilling to memory occurs. Code generation then emits assembly instructions, incorporating peephole optimizations to refine short sequences for better performance, such as combining multiple moves into a single instruction. Modern back ends, such as those in , support retargeting across multiple instruction set architectures (ISAs) through modular designs, allowing reuse of front and middle ends while customizing selection rules, scheduling, and emission for each target. For example, LLVM's TableGen tool defines instruction encodings and patterns, enabling back ends for over 20 ISAs including x86 and . The overall process proceeds from optimized to a DAG representation for selection, followed by to respect dependencies and hazards, , and final assembly emission into or binaries. A representative example is compiling a simple IR assignment like %result = add i32 %a, 5 on x86: instruction selection might choose add eax, 5 after loading %a into eax, with register allocation ensuring no conflicts via coloring. For register allocation, the greedy graph coloring algorithm approximates the chromatic number by iteratively assigning the smallest available color (register) to nodes ordered by decreasing degree, building the interference graph first where edges connect variables live simultaneously; this heuristic guarantees at most \Delta(G) + 1 colors, where \Delta(G) is the maximum degree, often sufficient for typical register counts like 16 on x86. Retargeting challenges arise from ISA differences, such as ARM's load-store architecture versus x86's register-memory operations, necessitating careful pattern matching to avoid inefficient code.

Correctness and Verification

Compiler correctness refers to the guarantee that the compiled preserves the semantics of the source program, ensuring equivalent observable behavior between the input and output across all possible executions. This semantic preservation is fundamental, as violations can lead to subtle where the generated produces incorrect results or crashes unexpectedly. Optimizing compilers prioritize while aiming to maintain this equivalence, whereas verifying compilers incorporate proofs to mathematically confirm it. Verification techniques for compilers span empirical testing and . Testing approaches include unit tests for individual components, integration tests for interactions, and tools like Csmith, which generates complex programs to expose discrepancies via differential testing against multiple compilers. Csmith has uncovered hundreds of bugs in production compilers such as and by producing inputs that trigger miscompilations. , in contrast, use theorem provers like to establish semantic preservation; the project exemplifies this by verifying a subset of to through a multi-pass , proving that optimizations do not alter . can also validate properties in specific phases, though it scales poorly to full compilers. Challenges in arise from the undecidability of checking semantic between arbitrary and target programs, stemming from the and requiring approximations or restrictions to subsets of languages. Phase errors, where transformations in one stage invalidate assumptions in another, further complicate guarantees. Examples include efforts in systems like , where the Mi-Cho-Coq framework uses to certify the functional correctness of Michelson smart contracts against their compiled execution semantics. Bug-finding tools like Csmith complement this by targeting undefined behaviors that evade formal proofs. Metrics for assessing verification effectiveness include test coverage rates, often aiming for over 90% coverage in regression suites to catch s, and self-hosting, where the compiler recompiles itself across multiple generations to validate consistency in output binaries. Self-hosting provides a practical , as discrepancies indicate bugs, though it does not prove full correctness. In recent developments, techniques such as large language models have been applied to detect missed optimizations and bugs in ; one approach identified 24 issues in production compilers by analyzing code for optimization opportunities. These methods, including IR-based crossover testing in , enhance traditional by generating targeted inputs that reveal subtle miscompilations. As of 2025, ongoing efforts include LLM-assisted tools like WhiteFox, which have detected additional bugs in .

Types of Compilers

Traditional Batch Compilers

Traditional batch compilers are systems that translate an entire program's source code into machine-executable form ahead of time, typically in an offline batch process without user interaction during compilation. This approach involves static analysis of the full codebase to generate optimized object code or binaries before program execution begins. Key characteristics include ahead-of-time (AOT) compilation, where the compiler performs all necessary analysis and optimization in a single or multi-pass run on the complete source, producing standalone executables or intermediate representations suitable for direct execution. Prominent examples are the GNU Compiler Collection (GCC), which compiles languages like C and C++ into native machine code, and the Java compiler (javac), which translates Java source files into platform-independent bytecode for the Java Virtual Machine. These compilers operate in batch mode, processing multiple input files as a unit to resolve dependencies and generate outputs like object files or class files. The typical workflow begins with source code input, which undergoes preprocessing (e.g., macro expansion in C), lexical analysis, parsing, semantic analysis, optimization, and code generation to produce object files containing relocatable machine code. These object files are then linked with libraries and resolved for external references to create a final executable binary, ready for immediate execution without further compilation. This process is managed via command-line invocation or build tools, ensuring reproducibility in non-interactive environments. Advantages of traditional batch compilers include the generation of highly optimized binaries that exhibit superior performance and minimal startup latency, as all overhead occurs upfront rather than during execution. They also eliminate runtime compilation costs, making them ideal for resource-constrained environments. However, disadvantages encompass extended times for large codebases, which can prolong development iterations, and a lack of adaptability to runtime conditions since optimizations are fixed at . Common use cases for batch compilers include , where native code efficiency is paramount, such as developing operating system kernels or device drivers with . They are also prevalent in development, where predictable performance and small footprint are essential, as AOT compilation avoids the overhead of just-in-time mechanisms in constrained hardware. The evolution of batch compilers traces back to the original compiler developed for the in the mid-1950s by and his team at , which introduced high-level language translation and optimization in a batch-oriented manner to simplify scientific . Over decades, this model advanced from single-language tools like early implementations to polyglot systems like , initiated in 1987, which now supports over a dozen languages and targets diverse architectures while retaining the core batch compilation paradigm.

Just-In-Time (JIT) Compilers

Just-in-time (JIT) compilers translate or (IR) into native during program execution, enabling dynamic optimization based on runtime behavior. This contrasts with by deferring code generation until necessary, often incorporating to target frequently executed code paths for improved efficiency. The JIT process typically starts with or baseline compilation for rapid startup, then progresses to full optimization as usage patterns emerge. The core mechanism of JIT compilers involves monitoring execution to identify "hot" code—regions executed often—and compiling them on-the-fly with aggressive optimizations unavailable at static compile time. For instance, tracing JITs record sequences of operations as linear traces, applying transformations like loop unrolling and dead code elimination to these traces for streamlined execution. In the V8 JavaScript engine, this pipeline includes an interpreter (Ignition) for bytecode generation, a baseline JIT (Sparkplug) for quick machine code, and advanced optimizing JITs (Maglev and TurboFan) that use static single assignment (SSA) form and runtime feedback on types and shapes to specialize code, with deoptimization handling failures in assumptions. Such approaches allow JITs to outperform pure interpretation while adapting to actual workloads. Prominent examples of JIT compilers include the JVM, released by in 1999, which uses a tiered system: the C1 client compiler for fast, lightweight optimization and the C2 server compiler for thorough analysis, significantly reducing interpretation overhead in Java applications. The .NET (CLR) applies JIT to convert (CIL) into platform-specific native code at method invocation, supporting managed execution across environments. , a trace-based JIT for the scripting language developed by Mike Pall since 2005, combines an assembler-based interpreter with advanced optimizations to achieve near-native speeds in dynamic scenarios like games and embedded systems, while maintaining compatibility with Lua 5.1. JIT compilers offer advantages in runtime adaptability, such as tailoring optimizations to observed branch probabilities and data types, which can yield higher peak performance than static methods for long-running or polymorphic code. This enables techniques like speculative inlining based on data, boosting throughput in applications with unpredictable paths. However, drawbacks include initial startup delays from on-the-fly , potentially increasing launch times by seconds in large programs, and elevated memory consumption from caching multiple code versions and profiles. Essential techniques in JIT compilers encompass adaptive optimization, which iteratively recompiles code using evolving runtime profiles to refine decisions like , and deoptimization, a mechanism to discard optimized code and resume from a interpreter if assumptions (e.g., values or type stability) invalidate. Deoptimization relies on inserted anchors—synchronization points that capture state—and assume checks to trigger on-stack replacement, preserving correctness in speculative scenarios like propagation or inlining. These methods, formalized in verified JIT frameworks, balance aggressive with reliable fallback, as seen in systems handling dynamic languages. In the 2020s, developments for in browsers have advanced compilation support, integrating , on-stack replacement, and recompilation to generate specialized native code efficiently while adhering to constraints. (PGO) has also matured in JIT environments, with .NET 9 expanding with Dynamic PGO to optimize more code patterns, such as type checks and casts.

Specialized Compilers

Specialized compilers are tailored for particular domains, architectures, or development needs, extending beyond general-purpose tools to address unique constraints such as target platform differences or specialized hardware. Cross-compilers, for instance, run on a host platform but produce executable code for a different target platform, enabling development for embedded systems or mobile devices without requiring the target hardware. The GNU Compiler Collection (GCC) supports cross-compilation through target-specific configurations, allowing developers to specify the host and target architectures via command-line options like --target=arm-linux-gnueabihf. A prominent example is the Android Native Development Kit (NDK), which uses cross-compilers to build C/C++ code for ARM-based Android devices from x86 hosts. Source-to-source compilers, also known as transpilers, translate code from one high-level language to another, often to ensure compatibility with existing environments or to leverage specific features. Babel transpiles modern (ES6+) to older versions (ES5) for broader browser support, using plugins to handle syntax transformations without altering semantics. Similarly, the TypeScript compiler (tsc) transpiles TypeScript—a superset of with static typing—to plain , enabling type-checked development while producing runnable code for web and environments. Domain-specific compilers optimize for niche or computational paradigms, incorporating specialized optimizations and backends. For graphics processing units (GPUs), NVIDIA's NVCC (CUDA Compiler Driver) compiles C++ code to parallel executable kernels for NVIDIA GPUs, handling device-specific instructions and through a phased compilation process that separates host and device code. In , Microsoft's Q# compiler translates quantum algorithms written in Q#—a —to executable operations for quantum simulators or like Quantum, supporting operations like entanglement and measurement. For , the XLA (Accelerated Linear Algebra) compiler optimizes and models by fusing operations into efficient kernels for CPUs, GPUs, or TPUs, reducing execution time through graph-level transformations without requiring code changes. Other specialized variants include incremental compilers, which recompile only modified program portions to accelerate iterative development, and bootstrapping compilers, which compile their own source code to enable self-hosting. Rust's compiler employs incremental compilation by tracking dependencies and caching intermediate results, allowing rapid rebuilds after small changes. Bootstrapping, as in GCC's build process, starts with a minimal compiler written to produce a full self-compiling version. Modern examples highlight practical applications: facilitates cross-compilation via rustup's target addition (e.g., rustup target add aarch64-unknown-linux-gnu), enabling seamless builds for multiple architectures including for systems. serves as a drop-in C/C++ compiler with built-in C interop, using @cImport to seamlessly link C headers and libraries while supporting cross-compilation out-of-the-box for targets like Windows from hosts. Emerging trends as of 2025 focus on compilers for devices, optimizing neural networks for low-power microcontrollers with techniques like quantization and operator fusion. Compiler-aware designs integrate -specific instructions into processors, enabling efficient on resource-constrained devices. These advancements support autonomous in without cloud reliance.

Advanced Topics

Optimization Techniques

Optimization techniques in compilers transform intermediate representations to improve program performance metrics such as execution time and code size while preserving semantics. These techniques are broadly classified into local optimizations, which operate on small code segments; global optimizations, which span basic blocks within a function; and interprocedural optimizations, which analyze across function boundaries to enable more aggressive transformations. Local optimizations, such as , examine short sequences of instructions—typically 1 to 4 machine words—and replace them with equivalent but more efficient code, such as eliminating redundant loads or jumps. This approach, first systematically explored in early compiler designs, can reduce instruction count in machine-specific phases. Global optimizations leverage across an entire to eliminate , propagate constants, or reorder computations for better register utilization. Interprocedural analysis extends this by constructing call graphs and summary information, allowing optimizations like inlining or alias elimination that cross boundaries, improving execution speed in large programs through reduced call overhead. Loop optimizations form a critical subset, targeting iterative structures common in numerical and data-processing code. replicates the loop body to reduce overhead from increment and test operations; for instance, the following C loop:
for (int i = 0; i < 4; i++) {
    a[i] = b[i] + c;
}
may be transformed into:
a[0] = b[0] + c;
a[1] = b[1] + c;
a[2] = b[2] + c;
a[3] = b[3] + c;
This eliminates branch instructions but increases code size, trading compactness for potential speedups on small fixed iterations by enabling better . Loop fusion merges adjacent loops with compatible iteration spaces to enhance data locality and efficiency, reducing accesses in array-based computations and improving performance in memory-bound workloads. Advanced techniques build on these foundations to exploit features. Vectorization, particularly using SIMD instructions, automatically converts scalar loops into vector operations that process multiple elements in parallel, as implemented in compilers like Intel's , which can achieve significant speedups on supported architectures for aligned array computations. Auto-parallelization identifies independent loop iterations and inserts threading directives, such as pragmas, to distribute work across cores without manual intervention, though effectiveness depends on loop dependence analysis and can deliver performance gains on multicore systems for data-parallel kernels. (PGO) uses runtime execution profiles from instrumented binaries to inform decisions like branch prediction or hot-path inlining, resulting in 5-15% average speedups across diverse workloads by aligning optimizations with actual usage patterns. Recent integrations of enhance traditional heuristics. In LLVM's MLGO framework, models, trained on optimization sequences from prior compilations, guide phase ordering to select pass orders that minimize execution time, demonstrating 1.3-2.5% geomean on the SPEC CPU2017 benchmark suite over default heuristics. Neural networks applied to , such as graph neural networks modeling graphs, learn spilling and coloring decisions to reduce live ranges and evictions. Performance is quantified using , defined as: \text{Speedup} = \frac{T_{\text{original}}}{T_{\text{optimized}}} where T denotes execution time; values above 1 indicate improvement, with SPEC benchmarks like CPU2006 or providing standardized metrics for compiler optimizations. Amdahl's law bounds parallel optimization potential: S = \frac{1}{f + \frac{(1-f)}{p}} where f is the fraction of sequential code (0 ≤ f ≤ 1) and p is the number of processors. This formula derives from total execution time T = f + \frac{1-f}{p} (normalizing original time to 1), highlighting that speedup approaches $1/f as p \to \infty, emphasizing the need to minimize sequential portions for scalable gains; for example, with f=0.05 and p=16, S \approx 9.1. As of 2025, further advancements include large language models (LLMs) for compiler optimization, enabling self-improving systems that iteratively refine heuristics based on performance feedback, and techniques like Iterative BC-Max for binary size reduction. These approaches show promise in achieving additional performance gains, such as up to 1.75x speedups in some cases using models like CodeLlama.

Parallel and Distributed Compilation

Parallel compilation techniques enable compilers to leverage multi-core processors by executing independent phases or tasks concurrently, significantly reducing overall build times for large codebases. For instance, the compiler driver supports parallel job execution through a built-in scheduler that compiles independent translation units simultaneously, allowing developers to specify the number of jobs via build system flags like -j in Make or . This approach is particularly effective for phases such as , , and , where dependencies between modules can be modeled to avoid conflicts. Build systems like employ dependency graphs to represent task interdependencies, enabling efficient parallel scheduling by default based on the system's CPU count, which ensures maximal utilization without manual intervention. Distributed compilation extends parallelism across networked machines, distributing workload to remote hosts for even greater scalability in team or CI environments. Tools like Bazel facilitate this through remote caching, where build actions—defined by inputs, outputs, and commands—are stored on a shared server; subsequent builds retrieve cached artifacts via HTTP, avoiding redundant computations and ensuring reproducible results across machines. Complementing this, ccache acts as a local or remote compiler cache for incremental builds, hashing source files and compiler arguments to reuse precompiled objects, which accelerates recompilations after minor changes by bypassing full preprocessing and compilation steps. In networked setups, DistCC distributes preprocessed C/C++ source code to volunteer machines running a daemon, compiling in parallel without requiring shared filesystems or identical environments, thus scaling builds nearly linearly for small clusters. Implementing and distributed compilation introduces challenges, including race conditions in shared data structures like symbol tables, where concurrent accesses during semantic or linking can lead to inconsistent results if not synchronized via locks or operations. Load balancing poses another issue in distributed systems, as uneven task distribution across heterogeneous nodes—due to varying CPU speeds or network latency—can cause idle resources and prolonged build times, necessitating dynamic scheduling algorithms to monitor and reassign workloads. Prominent tools include the Ninja build system, which optimizes parallel execution through its lightweight dependency resolution, achieving sub-second incremental builds for projects with tens of thousands of files like Chromium. In LLVM, parallel code generation during Link-Time Optimization (LTO) distributes backend tasks across threads, speeding up the transformation of intermediate representation (IR) to machine code for large modules. These techniques yield substantial benefits, such as reducing Chromium's full rebuild times from hours to under an hour on multi-core systems through jumbo builds that minimize inter-module dependencies, enabling higher parallelism.

References

  1. [1]
    What is a compiler? - Brown CS
    A compiler is a program that takes in source code written in one language (called the source language) and returns source code written in another language ...
  2. [2]
    [PDF] What does a compiler do?
    What's a Compiler??? At the very basic level a compiler translates a computer program from source code to some kind of executable code:.
  3. [3]
    Compiler Architecture
    Compiler Architecture · Overview · Compiler Components · Lexical Analysis (Scanning) · Syntax Analysis (Parsing) · Combining Scanning and Parsing · Semantic Analysis.
  4. [4]
    [PDF] History of Compilers - cs.wisc.edu
    The term 'compiler' was coined in the early 1950s by Grace Murray Hopper. The FORTRAN compiler of the late 1950s was one of the first real compilers.
  5. [5]
    Compilers and Program Analysis
    They play a critical role in optimizing the performance, efficiency, and security of applications by improving the way programs are written and executed on ...
  6. [6]
    [PDF] Compiler Research: The Next 50 Years
    Feb 2, 2009 · Compilers play a critical role in en- hancing computer security by reducing the occurrence of vulnerabilities due to coding errors and by ...
  7. [7]
    Compiler | Encyclopedia of Computer Science - ACM Digital Library
    A compiler is a program that translates programs expressed in a source language into equivalent programs expressed in a target language.
  8. [8]
    CS 4120 Spring 2023 Introduction to Compilers
    A compiler is a translator from one language, the source language to another, the target. Translation is not an easy task when the source and target languages ...
  9. [9]
    Intermediate Representation - Communications of the ACM
    Dec 1, 2013 · A compiler is a software program that translates a high-level source-language program into a form ready to execute on a computer. Early on in ...
  10. [10]
    What is a Compiler? | IBM
    Compilers are used to convert the types of code developers prefer to use into the types of code that systems require to operate. In this way, compilers improve ...What is a compiler? · How compilers work
  11. [11]
    What Is a Software Compiler? - Wind River Systems
    A software compiler is a program that translates source code written in a high-level programming language into machine code that can be executed by a computer.
  12. [12]
    What is a compiler? | Definition from TechTarget
    Apr 11, 2025 · A compiler is a special program that translates a programming language's source code into machine code, bytecode or another programming language.<|separator|>
  13. [13]
    [PDF] ThE STRucTuRE OF A COMpiLER - cs.wisc.edu
    Semantic analysis also begins the synthesis phase. The synthesis phase may translate source programs into some intermediate representation (IR) or it may.
  14. [14]
    Milestones:A-0 Compiler and Initial Development of Automatic ...
    Oct 4, 2024 · During 1951-1952, Grace Hopper invented the A-0 Compiler, a series of specifications that functioned as a linker/loader.
  15. [15]
    Did Grace Hopper Create the First Compiler?
    Dec 21, 2022 · A compiler method, according to Hopper's definition, was a program that copied the subroutine code into the proper place in the main program ...
  16. [16]
    [PDF] Lecture 27 C and Assembly int main(){ int x=10,y=15
    Using the GNU C compiler –S option, we can generate the assembly code for a source code. For example, consider the program simple.c int main(){ int x=10,y ...
  17. [17]
  18. [18]
    None
    ### Summary of Compiler Anatomy and Components
  19. [19]
  20. [20]
  21. [21]
  22. [22]
  23. [23]
    Lex - A Lexical Analyzer Generator by M. E. Lesk and E. Schmidt
    Lex is a program generator designed for lexical processing ofharacter input streams. It accepts a high-level, problem oriented specification for character ...Missing: Mike | Show results with:Mike
  24. [24]
    Yacc Yet Another Compiler Compiler by Stephen C. Johnson
    Yacc provides a general tool for describing the input to a computer program. The Yacc user specifies the structures of his input, together with code to be ...
  25. [25]
    Compiler vs. Interpreter: 12 Critical Differences To Know - Spiceworks
    Jun 16, 2023 · A compiler converts the whole source code to object code while an interpreter transforms and runs the source code line by line.
  26. [26]
    Compiler vs. Interpreter in Programming | Built In
    Compiled code runs faster, while interpreted code runs slower. A compiler displays all errors after compilation. If your code has mistakes, it will not compile.
  27. [27]
    4. Execution model — Python 3.14.0 documentation
    Each interpreter completely encapsulates all of the non-process-global, non-thread-specific state needed for the Python runtime to work. Notably, the ...
  28. [28]
    Interpreted vs Compiled Programming Languages - freeCodeCamp
    Jan 10, 2020 · Interpreted languages tend to be more flexible, and often offer features like dynamic typing and smaller program size. Also, because ...
  29. [29]
    Compiler vs. Interpreter (With Difference and Benefits) - Indeed
    Mar 28, 2025 · Benefits of using an interpreter: · It's easy to debug. Code debugging is easier with an interpreter because it reads the code line by line.Missing: disadvantages | Show results with:disadvantages
  30. [30]
    Understanding the Execution of Python Program - GeeksforGeeks
    Jul 10, 2020 · Python program execution involves two steps: compilation to bytecode, and then the interpreter converting bytecode to machine code. JIT can ...
  31. [31]
    The structure and performance of interpreters - ACM Digital Library
    Interpreted languages have become increasingly popular due to de- mands for rapid program development, ease of use, portability, and safety.<|separator|>
  32. [32]
    Compiler vs Interpreter: Key Differences in Program Execution ...
    Jun 26, 2025 · The most notable disadvantage is slower execution speed. Because code is translated and executed on-the-fly during every run, interpreted ...<|separator|>
  33. [33]
    [PDF] History of Lisp - John McCarthy
    Feb 12, 1979 · The implementation of LISP began in Fall 1958. The original idea was to produce a compiler, but this was considered a major undertaking, and we ...
  34. [34]
    7.4 Programming Language Implementation - OpenStax
    Nov 13, 2024 · As previously discussed, the major methods of implementing programming languages are compilation, pure interpretation, and hybrid implementation ...
  35. [35]
    Introduction of Assembler - GeeksforGeeks
    Jul 11, 2025 · An Assembler can be defined as a program that translates an assembly language program into a machine language program. Self-assembler is a ...
  36. [36]
    Linker - GeeksforGeeks
    Jul 11, 2025 · The linker's job is to manage and connect different pieces of code and data, ensuring that all references between them are properly resolved.
  37. [37]
  38. [38]
    LD
    Summary of each segment:
  39. [39]
    Compiler, Linker, Assembler, and Loader - Baeldung
    Mar 18, 2024 · So, the linker takes all object files as input, resolves all memory references, and finally merges these object files to make an executable file ...2. Executable Generation · 3. Compiler · 5. Linker
  40. [40]
    Kathleen Booth (1922 - 2022) - Biography - MacTutor
    Kathleen Booth was a pioneer in computer development being the first to create assembly language and, with her husband, produced the "Booth multiplier ...
  41. [41]
    [PDF] Programming in America in the 1950s -- Some Personal Impressions
    Some idea of the machine difficulties facing early programmers can be had by a brief survey of a few of the bizzare characteristics of the Selective. Sequence ...
  42. [42]
    The history of Fortran I, II, and III - ACM Digital Library
    The FORTRAN language described in the Programmer's Reference Manual dated Oc- tober 15, 1956 (IBM, 1956) differed in a few respects from that of the Preliminary ...
  43. [43]
    [PDF] The Early Development of Programming Languages. - DTIC
    L-rooker's Autocode. Back in b-land - a: 1 er, R. A. Brooker introduced a new t~~e of Autocode for the Mark I machine. This language -wa s much “cleaner ...
  44. [44]
    Fortran - IBM
    John Backus, Fortran's primary author, described the process as “hand-to-hand combat with the machine,” with the machine often winning. The cost of ...
  45. [45]
    Niklaus E. Wirth - A.M. Turing Award Laureate
    Algol 60, the most important creation of the Algol group, introduced recursive functions, structured code blocks, and local variables. It also pioneered the ...
  46. [46]
    Syntax-directed compiling | Proceedings of the April 21-23, 1964 ...
    This paper is primarily concerned with the analysis of source statements in a programming language, although some of the ideas and techniques may be ...
  47. [47]
    [PDF] FORTRAN, PROGRAMMING LANGUAGE - CERN Document Server
    the optimizing aid the FREQUENCY statement. A first version of FORTRAN IV for the IBM 7030 was released in 1962, and within a year other versions were ...
  48. [48]
    The Development of the C Language - Nokia
    In 1978 Brian Kernighan and I published The C Programming Language [Kernighan 78]. ... Dennis Ritchie turned B into C during 1971-73, keeping most of B's syntax ...
  49. [49]
    The Development of the C Language - Nokia
    This paper is about the development of the C programming language, the influences on it, and the conditions under which it was created.
  50. [50]
    [PDF] A Tour Through the Portable C Compiler - Semantic Scholar
    A portable compiler: theory and practice · Stephen C. Johnson. Computer Science. POPL. 1978. TLDR. An overview of the compiler structure and algorithms is given ...
  51. [51]
    Abstractions, Their Algorithms, and Their Compilers
    Feb 1, 2022 · Modern compilers refine the translation process into a composition of phases, where each phase translates one representation of the source ...
  52. [52]
    A compiler language for data structures - ACM Digital Library
    The third phase of the compiler is essen- tially a macro generator which uses the intermediate code file as macro calls and outputs assembly code. The macro ...<|control11|><|separator|>
  53. [53]
    (PDF) Compiler Construction - The Art of Niklaus Wirth.
    description of this technique can be found in [3]. 8 Single Pass versus Multi Pass Compilers. Wirth designed his languages so that they can be compiled in a ...
  54. [54]
    [PDF] The History of Fortran I, II, and III by John Backus
    It describes the development of the optimizing compiler for Fortran I, of various language manuals, and of Fortran II and III. It concludes with re- marks about ...
  55. [55]
    A survey of compiler optimization techniques - ACM Digital Library
    The history of optimizing compilers dates back at least as far as FORTRAN I (1). At that time, most programming was done in machine language, and a compiler ...
  56. [56]
    96 ASSEMBLER - ACM Digital Library
    The first pass through the source program collects all the symbol definitions into a symbol table, and the second pass converts the program to binary symbolic ...
  57. [57]
    A NELIAC-generated 7090-1401 compiler - ACM Digital Library
    Published: 01 February 1962 Publication History. 5citation303Downloads ... A basic “bootstrap” process was used to generate all but the first, i.e., the ...<|separator|>
  58. [58]
    Using the New Pass Manager — LLVM 22.0.0git documentation
    The optimization pipeline (aka the middle-end) uses the new PM, whereas the backend target-dependent code generation uses the legacy PM.
  59. [59]
    [PDF] CS153: Compilers Lecture 19: Optimization - Harvard University
    Optimization are code transformations: ... Additional optimizations for the specific arguments can be enabled (e.g., copy propagation, dead code elimination).
  60. [60]
    [PDF] Optimizations - UT Computer Science
    Control Flow Graphs. • Control Flow Graph (CFG) = graph representation of computation and control flow in the program. – framework for static analysis of ...
  61. [61]
    Reaching definitions and SSA - CS [45]12[01] Spring 2023
    Reaching definitions attempts to determine which definitions may reach each node in the CFG. A definition reaches a node if there is a path from the definition ...Missing: end | Show results with:end
  62. [62]
    [PDF] Chapter 7: - Loop optimisations Part 1: Reaching definitions
    Feb 21, 2022 · Reaching definitions – another data flow analysis. • Dataflow equations: ReachIn(n) = ReachOut(p). ReachOut(n) = Gen(n) U (ReachIn(n) – Kill(n)).Missing: middle end
  63. [63]
    [PDF] CS153: Compilers Lecture 20: Dataflow analysis - Harvard University
    •Liveness analysis is one example of dataflow analysis. •Other examples: Available Expressions, Reaching Definitions,. Constant-Propagation Analysis, … 9. Page ...
  64. [64]
    [PDF] Lecture 9: Loop Invariant Computation and Code Motion
    • Precise algorithm for code motion. • Use of reaching definitions and dominators in optimizations. 15-745: Loop Invariance. 13. Page 14. Carnegie Mellon. III ...Missing: optimization | Show results with:optimization
  65. [65]
    [PDF] CSC D70: Compiler Optimization LICM: Loop Invariant Code Motion
    The natural loop of a back edge is the smallest set of nodes that includes the head and tail of the back edge, and has no predecessors.
  66. [66]
    LLVM's Analysis and Transform Passes
    Hoisting operations out of loops is a canonicalization transform. It enables and simplifies subsequent optimizations in the middle-end.
  67. [67]
    LLVM: The middle-end optimization pipeline - nikic's Blog
    Apr 7, 2023 · The middle end performs optimizations that are independent (to the degree that this is possible) from both the input language and the output target.
  68. [68]
    Writing an LLVM Backend — LLVM 22.0.0git documentation
    This document describes techniques for writing compiler backends that convert the LLVM Intermediate Representation (IR) to code for a specified machine or ...
  69. [69]
    [PDF] BURG -- Fast Optimal Instruction Selection and Tree Parsing
    BURG is a program that generates a fast tree parser using BURS (Bottom-Up Rewrite System) technology. It accepts a cost-augmented tree grammar and emits a C ...Missing: back seminal paper<|separator|>
  70. [70]
    [PDF] Register Allocation via Graph Coloring by Preston Briggs
    Chaitin and his colleagues at IBM in Y or k town Heights built the fi rst global register allocator based on graph coloring. This thesis describes a series ...
  71. [71]
    [PDF] Instruction Selection - Department of Computer Science
    Apr 20, 2016 · In their 1994 paper, Ferdinand et al. demonstrate how finite tree automata can be used to solve both pattern matching and pattern selection—.
  72. [72]
    [PDF] Formal verification of a realistic compiler - Xavier Leroy
    An obviously better approach is to apply formal methods to the compiler itself in order to gain assurance that it pre- serves the semantics of the source ...
  73. [73]
    The CompCert verified compiler
    Sep 1, 2025 · CompCert is a compiler that generates ARM, PowerPC, RISC-V and x86 assembly code from CompCert C, a large subset of the C programming language.
  74. [74]
    [PDF] Finding and Understanding Bugs in C Compilers - Virtual Server List
    Mar 18, 2011 · Abstract. Compilers should be correct. To improve the quality of C compilers, we created Csmith, a randomized test-case generation tool, and.
  75. [75]
    Mi-Cho-Coq, a framework for certifying Tezos Smart Contracts - arXiv
    Sep 18, 2019 · In this article, we present Mi-Cho-Coq, a Coq framework for verifying the functional correctness of Michelson smart contracts.Missing: compiler | Show results with:compiler
  76. [76]
    [PDF] FLUX: Finding Bugs with LLVM IR Based Unit Test Crossovers
    Sep 1, 2023 · To find compiler bugs, previous work has employed fuzzing to automatically generate bug-triggering test cases. Unlike general-purpose software ...
  77. [77]
    What is a batch compiler? - Computer Science Stack Exchange
    May 30, 2017 · A JIT (Just-In-Time) compiler compiles code at run-time, ie as the program is running. Therefore the cost of compilation is part of the execution time of the ...Missing: traditional characteristics workflow examples
  78. [78]
  79. [79]
    The javac Command
    ### Summary of javac as a Batch Compiler
  80. [80]
    [PDF] Compiling Java for Embedded Systems
    Specifically, an optimizing, ahead-of-time compiler allows much better optimization along with much faster application start-up times than with JIT translators.Missing: disadvantages | Show results with:disadvantages
  81. [81]
    What are the advantages of just-in-time compilation versus ahead-of ...
    Jan 21, 2010 · A better JIT-compiler will improve the performance of existing programs. The Java code you wrote ten years ago will run faster today. Adapting ...What's the difference between ahead of time compiled code and ...c# - When does ahead-of-time (AOT) compilation happen?More results from stackoverflow.comMissing: traditional javac
  82. [82]
    Java Compilers for Embedded Systems - Barr Group
    Sep 1, 1998 · Java compilers will allow developers to precompile their Java code, much the same as traditional compiled languages, such as C or C++. Besides ...
  83. [83]
    History of FORTRAN and FORTRAN II - Software Preservation Group
    The goal of this project is to preserve source code, design documents, and other materials concerning the original IBM 704 FORTRAN/FORTRAN II compiler.<|control11|><|separator|>
  84. [84]
    The JIT compiler - IBM
    The Just-In-Time (JIT) compiler is a component of the runtime environment that improves the performance of Java™ applications by compiling bytecodes to native ...
  85. [85]
    Managed Execution Process - .NET - Microsoft Learn
    Apr 20, 2024 · Compiling CIL to native code. At execution time, a just-in-time (JIT) compiler translates the CIL into native code.Missing: mechanism | Show results with:mechanism
  86. [86]
    Maglev - V8's Fastest Optimizing JIT - V8 JavaScript engine
    Dec 5, 2023 · Maglev sits between our existing Sparkplug and TurboFan compilers, and fills the role of a fast optimizing compiler that generates good enough code, fast ...
  87. [87]
    Surgical precision JIT compilers - ACM Digital Library
    Just-in-time (JIT) compilation of running programs provides more optimization opportunities than offline compilation. Modern JIT compilers, such as those in ...
  88. [88]
    [PDF] A Brief History of Just-In-Time - Department of Computer Science
    Software systems have been using “just-in-time” compilation (JIT) techniques since the. 1960s. Broadly, JIT compilation includes any translation performed ...
  89. [89]
    LuaJIT is a Just-In-Time Compiler (JIT) for the Lua programming ...
    Overview. LuaJIT has been successfully used as a scripting middleware in games, appliances, network and graphics apps, numerical simulations, trading ...
  90. [90]
    PAYJIT: space-optimal JIT compilation and its practical implementation
    Just-in-time (JIT) compilation in dynamic programming languages can improve execution speed in code with hot sections. However, that comes at the cost of ...Missing: definition | Show results with:definition
  91. [91]
    Formally Verified Speculation and Deoptimization in a JIT Compiler
    Just-in-time compilers for dynamic languages routinely generate code under assumptions that may be in- validated at run-time, this allows for specialization ...
  92. [92]
    What's new in .NET 9 runtime - Microsoft Learn
    Nov 11, 2024 · .NET 8 enabled dynamic profile-guided optimization (PGO) by default. NET 9 expands the JIT compiler's PGO implementation to profile more code ...<|separator|>
  93. [93]
    Fifty years of peephole optimization - Semantic Scholar
    Peephole optimization is covered in most textbooks on compiler construction, has been used in many research-purpose and production-quality compilers, ...Missing: seminal | Show results with:seminal
  94. [94]
    [PDF] Lecture Notes on Loop Optimizations
    Mar 26, 2024 · Unrolling Sometimes it is beneficial to place a copy of the body of the loop in a new code block in front of the loop. This can lead to new ...
  95. [95]
    [PDF] a guide to vectorization with intel® c++ compilers
    This Guide will focus on using the Intel® Compiler to automatically generate SIMD code, a feature which will be referred as auto-vectorization henceforth. We ...
  96. [96]
    Auto-Parallelization and Auto-Vectorization | Microsoft Learn
    Oct 17, 2022 · Auto-Parallelizer and Auto-Vectorizer are designed to provide automatic performance gains for loops in your code.
  97. [97]
    [PDF] MLGO: a Machine Learning Guided Compiler Optimizations ... - arXiv
    The paper is organized as follows: Section 2 provides an overview of the relevant ML techniques and the LLVM in- liner. Section 3 gives an overview of MLGO: our ...Missing: phase | Show results with:phase
  98. [98]
    [PDF] REGISTER ALLOCATION USING GRAPH NEURAL NETWORKS
    In this pa- per we propose a solution for register allocation problem of compiler using Deep. Learning networks based on graphs called as graph neural networks ...
  99. [99]
    SPEC CPU®2017 Overview / What's New?
    Sep 3, 2024 · SPEC CPU 2017 provides a comparative measure of integer and/or floating point compute intensive performance. If this matches with the type of ...
  100. [100]
    [PDF] Validity of the Single Processor Approach to Achieving Large Scale ...
    Amdahl. TECHNICAL LITERATURE. This article was the first publica- tion by Gene Amdahl on what became known as Amdahl's Law. Interestingly, it has no equations.
  101. [101]
    D69582 Let clang driver support parallel jobs
    Oct 29, 2019 · This patch implements a simple scheduler which let clang driver compile independent jobs in parallel.Oct 29 2019, 1:26 PM · Oct 29 2019, 2:38 PM
  102. [102]
    The Ninja build system
    Builds are always run in parallel, based by default on the number of CPUs your system has. Underspecified build dependencies will result in incorrect builds.Missing: IR | Show results with:IR
  103. [103]
    Remote Caching | Bazel
    This page covers remote caching, setting up a server to host the cache, and running builds using the remote cache.
  104. [104]
    Ccache — Compiler cache
    ### Summary of ccache for Incremental Builds
  105. [105]
    distcc: a fast, free distributed C/C++ compiler
    ### Summary of DistCC for Networked Compilation Clusters
  106. [106]
    Load Balancing Approach in Distributed System - GeeksforGeeks
    May 13, 2024 · A load balancer is a device that acts as a reverse proxy and distributes network or application traffic across a number of servers.
  107. [107]
    Big Project Build Times–Chromium | Random ASCII - WordPress.com
    Mar 30, 2020 · These jumbo builds significantly reduced the time to do full rebuilds of Chromium on machines with few processors. For incremental builds and ...