Fact-checked by Grok 2 weeks ago

Failure cause

In and reliability contexts, a failure cause refers to the specific circumstances or factors during the design, manufacture, or operational use of an item that lead to its , defined as the termination of the ability to perform a required . This concept is central to , where identifying the root cause enables engineers to mitigate risks, enhance product durability, and prevent similar incidents in complex systems such as bridges, , or electronic devices. Failure causes manifest across the of engineering artifacts, from initial flaws—such as inadequate load calculations in structural components—to defects like improper material processing, and usage-related issues including environmental exposure or misuse. For instance, the 1907 collapse stemmed from errors in strut dimensions and load assumptions, while inconsistencies have caused cracks in components like bimetal bearings due to suboptimal . Even post-operational stages, such as retirement, can reveal causes related to environmental impacts, like toxic materials hindering safe disposal. These causes are classified at item level (affecting individual units) or type level (impacting an entire product line, as in automotive recalls). Key categories of failure causes include mechanical, material, process-related, and human-induced factors, each contributing to distinct degradation paths. Mechanical causes, such as from overload, due to cyclic stresses, or from , often arise in rotating machinery like pumps or turbines. Material causes involve inherent weaknesses, like in metals exposed to , leading to structural weakening. Process-related causes stem from production errors, such as during , while human-induced ones encompass errors in or , exemplified by misalignment in causing leaks. Other mechanisms like from fluid flow or under sustained high temperatures further illustrate how environmental and operational stresses precipitate . Understanding and addressing failure causes is essential for , employing tools like Failure Modes and Effects Analysis (FMEA) to systematically evaluate potential causes, their effects, and criticality. This proactive approach not only reduces safety risks and economic losses but also informs design improvements, as seen in NASA's analyses of common cause failures in space systems to achieve ultra-high reliability. By tracing causes through laboratory examinations, such as metallurgical testing, engineers can isolate root factors like defects or overloads, ensuring robust performance across industries.

Definitions and Terminology

Definition of Failure Cause

In and reliability contexts, a failure cause refers to the underlying circumstances during , manufacture, or use that lead to the malfunction or breakdown of a component, , or process. According to the International Electrotechnical Vocabulary (IEV 191-04-17), failure cause is defined as "the circumstances during , manufacture, or use which have led to a ." These causes are often rooted in initiating factors such as defects, inadequate , or external stressors like excessive loads or environmental conditions, which disrupt intended functionality. Identifying these root conditions is essential for tracing how they propagate to observable failure modes, where the cause initiates the breakdown process. The importance of pinpointing failure causes lies in its central role within , where it enables proactive measures to enhance system performance and longevity. By systematically analyzing causes, engineers can reduce unplanned downtime, which accounts for significant productivity losses in industrial settings, often cutting interruptions by up to 45% through targeted interventions. Similarly, understanding these causes minimizes costs by preventing recurring defects in products, with reliability practices shown to lower claim expenses through improved and testing protocols. Moreover, it mitigates risks associated with catastrophic breakdowns, such as those in high-stakes environments like chemical processing or , thereby protecting personnel and . This formalized study of failure causes emerged prominently in the early , spurred by devastating industrial accidents that exposed vulnerabilities in practices. Events like the 1907 collapse, attributed to flaws including miscalculated loads leading to of structural chords, highlighted the dire consequences of overlooked causes and prompted the development of rigorous methods to prevent future incidents. For instance, a structural component subjected to unintended overload might deform and collapse, illustrating how a single initiating factor can into total system failure if not anticipated during .

Distinction from Failure Mode and Mechanism

In , a failure mode describes the observable manner or effect in which a or component ceases to function as intended, such as a structural or fluid leakage. This term focuses on the end-state symptom of , providing a description of the deviated performance or condition. A mechanism, by contrast, refers to the underlying physical, chemical, electrical, or other process that physically enables or drives the mode to occur, such as crack propagation resulting from loading. Mechanisms represent the intermediate deterioration processes that degrade material properties or system integrity over time. The cause constitutes the root initiator of the sequence, typically an originating factor like a design flaw that permits excessive cyclic or a decision to omit protective coatings, which allows the to initiate. Unlike modes and mechanisms, causes are often preventable through proactive measures and lie at the origin of the chain. These concepts form a hierarchical model: the cause triggers the , which culminates in the mode. This relationship can be represented as a simple —Failure Cause (e.g., inadequate ) → Failure (e.g., crack growth) → Failure Mode (e.g., component )—emphasizing the causal progression from origin to observable outcome. The distinctions are critical for : causes are the actionable origins addressable by or interventions (e.g., poor leading to susceptibility), modes serve as diagnostic symptoms, and mechanisms act as the bridging processes that can be modeled for prediction. In quantitative terms, rates influenced by such causes factor into reliability assessments, as seen in the reliability : R(t) = e^{-\lambda t} where R(t) is the probability of no failure by time t, and \lambda is the constant failure rate incorporating cause-related vulnerabilities.

Types of Failure Causes

Mechanical Causes

Mechanical causes of failure arise from the application of physical forces and stresses that exceed the of materials or structures, leading to deformation, , or loss of functionality. These failures occur when external loads surpass the inherent limits of a material's properties, often without involvement of . In contexts, such causes are analyzed through the lens of stress analysis and material under load. The primary mechanical causes include overload, fatigue, creep, and wear, each manifesting when stresses exceed limits in distinct ways. Overload refers to a sudden, excessive application of that surpasses the 's strength, resulting in immediate without prior damage accumulation. This typically happens in a single event, such as when an applied load exceeds the ultimate strength due to unexpected impacts or design oversights, producing clean, brittle surfaces. Fatigue failure develops progressively under repeated cyclic loading, even at stress levels below the material's yield strength, leading to crack initiation and until catastrophic . This involves the accumulation of microscopic over numerous cycles, often characterized by striations or beach marks on the fracture surface. is particularly prevalent in components subjected to fluctuating es, like rotating machinery or structural elements under traffic loads. The fatigue life N_f can be estimated using Basquin's , expressed as N_f = C (\Delta \sigma)^b, where \Delta \sigma is the stress , C is a material constant, and b is the fatigue strength exponent, typically ranging from -0.05 to -0.12 for metals. This empirical relation, derived from endurance tests, highlights the inverse power-law relationship between stress range and cycles to . Creep involves time-dependent plastic deformation under sustained constant , especially at elevated temperatures, where materials gradually elongate until rupture. It progresses through stages of primary deceleration, secondary steady-state , and tertiary acceleration with necking and void formation, ultimately causing failure when the deformation rate becomes unstable. Creep is critical in high-temperature applications like blades, where even moderate stresses lead to dimensional changes over time. Wear, another key cause, results from the progressive removal or deformation of from surfaces in under relative motion, reducing thickness and load-bearing capacity until failure by overload or ensues. Mechanisms include from hard particles, from surface under , and surface from repeated contacts, all diminishing the functional integrity of components like bearings or gears. Understanding these causes requires knowledge of fundamental material science concepts, such as yield strength, , and . Yield strength is the stress at which a material begins to deform plastically, marking the transition from elastic to permanent deformation, often determined using a 0.2% offset method on the stress-strain curve for materials without a clear yield point. represents the maximum stress a material can withstand before fracturing, corresponding to the peak of the engineering . The itself illustrates a material's response to uniaxial loading, starting with a linear elastic region (Hooke's law, where stress \sigma is proportional to strain \epsilon via modulus E: \sigma = E \epsilon), followed by yielding, plastic deformation, necking, and ultimate failure, providing a graphical basis for predicting mechanical limits. Representative examples of mechanical failures include the 1980 I-35W bridge inspection issues highlighting in plates, though the 2007 collapse involved flaws; pure mechanical overload cases, such as crane boom failures from excessive loads in controlled environments, demonstrate immediate without . These incidents underscore how mechanical stresses from loading can exceed limits in .

Corrosion and Chemical Causes

Corrosion and chemical causes refer to the of , primarily metals, through chemical reactions with their , leading to structural weakening and eventual . These processes are fundamentally electrochemical, involving the oxidation of the and of species in the surrounding medium, often resulting in the loss of integrity over time. Unlike purely mechanical , chemical proceeds continuously in the presence of suitable conditions, such as an , and can compromise components in , vehicles, and machinery. Key mechanisms of corrosion include uniform corrosion, where the material surface degrades evenly; pitting corrosion, which forms localized deep cavities; galvanic corrosion, occurring between dissimilar metals in electrical contact; and stress corrosion cracking, a combined chemical and tensile stress effect that propagates cracks. Uniform corrosion is characterized by a general thinning of the metal surface due to anodic dissolution across the entire exposed area, often seen in carbon steel pipelines exposed to moist soil. Pitting initiates at surface defects or inclusions, accelerating inward and potentially leading to perforation, as observed in stainless steels under chloride exposure. Galvanic corrosion arises when two metals with different electrochemical potentials are coupled in an electrolyte, with the more anodic metal corroding preferentially; for instance, aluminum alloys in aircraft structures can suffer accelerated attack when fastened to stainless steel components. Stress corrosion cracking involves the synergistic action of a corrosive environment and residual or applied tensile stress, resulting in brittle fractures; this has been a significant issue in buried natural gas pipelines, where near-neutral pH groundwater promotes crack growth. These mechanisms are influenced by factors such as pH, which affects the solubility of corrosion products—acidic conditions (low pH) generally accelerate dissolution—temperature, which increases reaction kinetics per the Arrhenius relation, and the presence of electrolytes that facilitate ion transport. The electrochemical principles underlying these processes are governed by the , which relates the to the concentrations of species involved in the half-cell reaction: E = E^0 - \frac{RT}{nF} \ln Q Here, E is the cell potential, E^0 is the standard potential, R is the , T is temperature in , n is the number of electrons transferred, F is Faraday's constant, and Q is the . In contexts, this equation predicts how shifts in environmental conditions, such as ion concentrations or , alter the driving force for anodic and cathodic reactions; for example, increased levels in coastal environments can lower the pitting potential of aluminum, promoting localized attack in aircraft fuselages. Rusting of steel pipelines exemplifies uniform driven by oxygen reduction as the cathodic reaction paired with iron oxidation, where the illustrates how higher temperatures elevate the corrosion rate by enhancing T and thus the logarithmic term. Environmental factors initiate and exacerbate these chemical causes, with exposure to moisture providing the necessary for conduction, salts like acting as aggressive anions that break down passive films on metals, and pollutants such as forming acidic solutions that lower pH and accelerate uniform attack. In marine or de-icing salt environments, ions penetrate layers on aluminum components, leading to pitting and subsequent structural degradation. Similarly, industrial pollutants in urban settings can deposit on exteriors, combining with rainwater to create electrolytic conditions that drive galvanic couples between the pipe coating and soil. These factors highlight the role of atmospheric and soil chemistry in failure causation, often requiring and protective measures to mitigate risks. can also interact with mechanical stresses, such as accelerating crack growth in load-bearing parts, though the primary degradation remains chemical. For example, the 2021 Surfside condominium collapse in involved of steel reinforcements contributing to structural failure.

Electrical and Thermal Causes

Electrical failures in components often stem from disruptions in or voltage integrity, leading to catastrophic damage. Short circuits occur when an unintended low-impedance path forms between conductors, allowing excessive to and generating intense that can melt components or ignite fires. Arcing, a related phenomenon, arises from the rapid breakdown of air or under , creating a channel that erodes contacts and produces explosive energy release. breakdown, frequently triggered by voltage surges or material aging, compromises the barrier, enabling unintended conduction and escalating to full . Thermal causes exacerbate these electrical issues by inducing stresses from temperature extremes. Overheating typically results from inadequate heat dissipation in densely packed or high-power circuits, where accumulated energy degrades materials and reduces operational lifespan. mismatches between dissimilar materials, such as in soldered joints or layered semiconductors, generate mechanical stresses during temperature cycling, potentially causing cracks or . A key principle underlying generation in resistive elements is Joule's law, which quantifies power as P = I^2 R, where P is power, I is , and R is ; this quadratic relationship with current amplifies thermal risks under overload conditions. In power transformers, overloads beyond rated capacity cause windings to overheat, accelerating degradation and risking explosive failure. Similarly, semiconductors like MOSFETs are prone to burnout from , where localized heating increases leakage current, further elevating temperature until junction meltdown occurs. can briefly accelerate electrical failures by degrading insulation integrity, though it primarily interacts through environmental exposure. Recent examples include incidents in electric vehicles, such as the 2023 recalls for lithium-ion overheating risks.

Human and Organizational Causes

Human and organizational causes of failure encompass errors stemming from individual actions, processes, and broader systemic issues within organizations that lead to unintended system breakdowns. These causes often arise from deviations in or institutional practices that compromise and reliability in complex systems, such as industrial plants, , and facilities. Unlike purely technical failures, these factors highlight the interplay between and processes, where lapses in judgment or oversight can initiate or amplify cascading events. Operator errors represent a primary category of human-induced failures, frequently resulting from inattention, procedural violations, or misjudgments during routine operations. For instance, in power plant settings, such errors have been linked to inadequate adherence to protocols, contributing to malfunctions and broader unreliability. Inadequate exacerbates these issues, as personnel lacking comprehensive preparation may overlook critical steps or fail to recognize hazards, a factor evident in historical incidents where insufficient skill development led to operational oversights. Poor schedules further compound risks, as deferred or improperly timed inspections allow minor defects to evolve into major failures; studies in aviation underscore how irregular servicing routines directly correlate with error-prone interventions that degrade component integrity. Design oversights due to insufficient testing constitute another key organizational cause, where assumptions during the phase bypass rigorous validation, embedding vulnerabilities that manifest under real-world stresses. Organizational culture profoundly influences these causes by shaping attitudes toward protocols; cultures prioritizing production over precaution often erode adherence to standards, fostering an environment where shortcuts become normalized and of near-misses is discouraged. Psychological factors, including cognitive biases, play a significant role in perpetuating errors—for example, during safety inspections leads operators to selectively interpret data in ways that affirm preconceived notions, ignoring contradictory that could prevent failures. Studies indicate that human factors contribute to approximately 70-90% of incidents in industries such as and , though exact figures vary by sector and methodology. A stark illustration is the 1986 nuclear disaster, where operator mistakes during a safety test—compounded by inadequate and flawed procedures—triggered a reactor explosion, highlighting how human errors can interact with design flaws to exacerbate physical issues. More recently, the 2010 involved organizational failures in oversight and , leading to environmental catastrophe. Addressing these causes requires targeted interventions, such as enhanced programs and cultural reforms, to mitigate their pervasive impact on system reliability.

Failure Scenarios and Analysis

Common Failure Scenarios

Common failure scenarios in engineering systems often follow recognizable archetypes that illustrate how initial anomalies escalate into complete breakdowns. One prevalent archetype is the , where the malfunction of a single component imposes additional on interconnected elements, propagating the issue across the system. For instance, in a mechanical assembly, an overloaded bearing may generate excessive and , which in turn accelerates on adjacent gears and shafts, leading to a of component failures. This sequence highlights the interconnected nature of systems, where localized overload—often in origin—can trigger widespread disruption. Another fundamental archetype is the curve , which models the temporal of s in populations of components or systems, particularly in mechanical and electronic applications. The , denoted as λ(t), varies over the lifecycle in a "bathtub" shape: a high initial decrease during the phase, a stable constant during useful life, and a subsequent rise in the wear-out phase. This arises from empirical observations in , where early failures stem from manufacturing defects, mid-life incidents are random, and late-stage breakdowns result from material degradation. In the infant mortality phase, λ(t) decreases rapidly as defective units are weeded out, often modeled using a Weibull distribution with shape parameter α < 1: \lambda(t) = \alpha \lambda^\alpha t^{\alpha - 1} where λ is the scale parameter and t is time. During the useful life phase, failures occur at a constant rate due to extrinsic factors, with λ(t) = constant, equivalent to an exponential distribution (α = 1). In the wear-out phase, λ(t) increases as aging effects dominate, again following a Weibull form but with α > 1, emphasizing the need to monitor systems beyond their stable period. These phases provide a conceptual framework for anticipating failure timing in design and operation. A generic narrative of such escalation appears in automotive systems, where neglected allows minor issues to compound into catastrophic . Over time, insufficient from skipped oil changes reduces protection, causing pistons and cylinders to overheat and expand, which binds and halts the entirely. This sequence underscores how routine , if unaddressed, transitions from subtle performance degradation to total operational failure.

Root Cause Analysis Methods

Root cause analysis (RCA) methods provide structured frameworks for dissecting failure causes after an incident, using failure scenarios as inputs to trace underlying factors systematically. (FTA), a key technique, originated in the 1960s at Bell Telephone Laboratories for the Minuteman missile project and was refined by for . Today, RCA methods integrate qualitative and quantitative approaches to move beyond surface-level symptoms. A typical step-by-step in begins with , involving gathering incident reports, logs, witness accounts, and physical evidence to establish a factual . This is followed by testing, where potential causes are formulated based on collected and evaluated through logical or statistical methods to identify the most probable contributors. Validation then occurs via techniques, such as simulations or cross-referencing with historical , to confirm the root cause and rule out alternatives. One foundational qualitative method is the 5 Whys technique, which iteratively asks "why" a problem occurred, typically five times, to peel back layers until reaching the root cause. Developed by in the early 1930s and popularized by at in the , it promotes simple, team-based questioning without requiring specialized tools, making it effective for operational failures like defects. The Fishbone diagram, also known as the , visually categorizes potential causes into branches representing factors such as materials, methods, machinery, and manpower, facilitating brainstorming sessions to map relationships to the failure effect. Invented by in 1968 for in , it emphasizes a holistic view of contributing elements, often used in cross-functional teams to uncover interconnected causes. For , () models failure pathways using a top-down with logic gates to represent event combinations leading to an undesired top event. Originating in 1962 at Bell Telephone Laboratories for the Minuteman missile project and refined by in the 1960s for , employs AND gates (requiring all inputs to occur) and s (requiring any input) to compute failure probabilities. The probability of the top event T for an OR gate with independent events is: P(T) = 1 - \prod_{i=1}^{n} (1 - P_i) where P_i are the probabilities of the input events, while for an AND gate it is: P(T) = \prod_{i=1}^{n} P_i This allows precise quantification of rare events in complex systems like nuclear or aviation safety. Software tools enhance FTA modeling by automating gate logic and probability calculations; for instance, ReliaSoft BlockSim supports fault tree construction with advanced features like binary decision diagrams for efficient analysis of large systems.

Prevention and Mitigation

Design Strategies

Design strategies in aim to proactively eliminate or minimize potential causes by incorporating reliability principles from the outset of the product development process. A fundamental principle is the (FOS), defined as the ratio of a material's ultimate strength to the allowable under operating conditions, typically expressed as \text{FOS} = \frac{\text{ultimate strength}}{\text{allowable stress}}. This approach accounts for uncertainties in material properties, loading conditions, and variations, ensuring structures can withstand loads beyond expected levels without ; for instance, common FOS values range from 1.5 to 4 depending on the application and risk level. Redundancy involves duplicating critical components or pathways to provide backups, thereby enhancing system reliability by allowing continued operation if a primary fails. In reliable , this can take forms such as active , where backups actively monitor and switch seamlessly, or passive , which activates only upon detection, reducing the overall probability of system . complements this by dividing systems into independent, interchangeable units that isolate causes, preventing localized issues—like a faulty subsystem—from propagating to the entire structure and facilitating easier replacement or upgrades. Key techniques include finite element analysis (FEA), a that simulates distributions and deformation in complex geometries to predict potential points before prototyping. By dividing a structure into finite elements and solving governing equations, FEA identifies high- regions under various loads, enabling designers to reinforce vulnerable areas and avert mechanical or thermal . further mitigates risks by intentionally operating components below their maximum specifications, such as limiting voltage to 80% of rated capacity or temperature to 70% of maximum, which reduces and extends component lifespan in electronics and mechanical systems. Industry standards guide these strategies to ensure consistent reliability across sectors. The standard, specifically tailored for automotive electrical and electronic systems, mandates hazard analysis and during design to achieve functional safety levels (ASILs), incorporating and to prevent malfunctions that could lead to hazardous events. Similarly, MIL-HDBK-338 provides military-specific guidelines for electronic reliability design, emphasizing FEA, redundancy allocation, and derating to meet stringent environmental and operational demands in defense applications.

Operational and Maintenance Practices

Operational and maintenance practices play a crucial role in detecting and mitigating failure causes during the active lifecycle of systems, shifting focus from initial to ongoing vigilance. techniques, such as vibration analysis and , enable early identification of anomalies like imbalances or overheating in machinery, allowing interventions before failures escalate. For instance, vibration monitoring uses sensors to detect frequency shifts indicative of wear, while infrared identifies thermal hotspots in electrical components. Recent advancements include integration of (AI) and for enhanced anomaly prediction and real-time analytics, as of 2025. These methods have been shown to reduce unplanned by up to 50% in industrial settings. Scheduled inspections form another cornerstone, often calibrated using (MTBF), defined as MTBF = 1/λ where λ represents the , to determine optimal intervals for checks. This approach ensures that components like bearings or are examined proactively based on historical reliability data, preventing degradation from progressing undetected. In and , such timetabled routines have extended equipment life by 20-30% through timely and alignment adjustments. Condition-based maintenance (CBM) relies on real-time monitoring to trigger actions only when predefined thresholds are exceeded, optimizing . Oil analysis, for example, assesses contamination and particulate levels to forecast failures in engines or gearboxes, correlating changes with impending . Thresholds might include particle counts above 10,000 per milliliter or iron content exceeding 100 , prompting immediate servicing. Studies in plants demonstrate that CBM can cut costs by 25% compared to time-based strategies. Integrating these practices organizationally involves comprehensive programs that equip personnel with skills in interpretation and adherence, fostering a culture of proactive . Feedback loops from past failures, such as post-incident reviews, refine procedures and update monitoring thresholds, ensuring continuous improvement. In facilities, such loops have significantly reduced human-error-related incidents.

References

  1. [1]
    Failure of Engineering Artifacts: A Life Cycle Approach - PMC
    Failure cause is defined as “the circumstances during design, manufacture or use which have led to a failure”. The analysis by IEC vocabulary stops here ...
  2. [2]
    [PDF] Chapter 3 Failures and Failure Classification - NTNU
    A failure is the termination of an item's ability to perform a required function, occurring when the function cannot be performed or falls outside the ...
  3. [3]
    Understanding Failure Mechanisms and Components | Reliability
    Apr 23, 2024 · The four failure mechanisms are mechanical (fracture, fatigue, corrosion), material, process-related, and human-induced failures.
  4. [4]
    None
    ### Summary of Failure Modes, Mechanisms, and Causes in Engineering
  5. [5]
    Failure Modes & Effects Analysis (FMEA) and Failure Modes ... - DAU
    FMEA/FMECA is a reliability technique examining potential failure modes and their effects on system performance, identifying ways systems can fail.<|control11|><|separator|>
  6. [6]
    [PDF] Common cause failures and ultra reliability 2
    They define a common cause failure as a “dependent failure in which two or more component fault states exist simultaneously, or within a short time interval, ...
  7. [7]
    Failure Analysis Services - Acuren
    Some common failure modes are fatigue, tensile overload, and stress rupture. Failure cause is determined from laboratory studies and knowledge of the component, ...
  8. [8]
    The Meaning of a Failure - Accendo Reliability
    A single failure in a complex system indicates that something is not right somewhere in the design or manufacturing process. It is difficult to know from a ...
  9. [9]
    How Can You Prevent Equipment Failure? Common Causes + ...
    6 Benefits Of Preventive Maintenance · Fewer Breakdowns And Work Interruptions · Decreased Downtime · Enhanced Efficiency · Longer Life Of Assets · Improved Safety.<|separator|>
  10. [10]
    Reliability Engineering | Quality-One
    Reliability Engineering, when implemented properly, can work to predict and prevent product failures thereby reducing warranty costs and increasing a business's ...
  11. [11]
    A smarter way to digitize maintenance and reliability - McKinsey
    Apr 23, 2021 · In addition, these programs can increase the useful life of high-cost components and reduce safety risks related to machine failure events.
  12. [12]
    Ten Engineering Disasters of the Last 100 Years - ASME
    Aug 17, 2023 · It is important for engineers to learn from failure. Here are 10 engineering failures that shook the industry, some from early in the 20th centurry.
  13. [13]
    What are Failure Modes? - Reliability Academy
    Oct 5, 2022 · Confused by failure terms? Learn the difference between failures, functional failures, failure modes, and mechanisms, explained simply and ...
  14. [14]
    [PDF] Failure Modes and Mechanisms - NERC
    A Failure Mechanism is the catalyst and physical path to failure. ▫ Equipment Failures have logical cause-and-effect relationships behind them. ▫ Physical ...
  15. [15]
    [PDF] Failure Analysis and Prevention: Fundamental causes of failure
    This chapter defines the failure and elaborates the conditions for failure of mechanical components. Further, the fundamental causes of failure associated with ...
  16. [16]
    Mechanical Overload - an overview | ScienceDirect Topics
    The failure or fracture of a product or component in a single event is known as mechanical overload. It is a common failure mode and may be contrasted with ...
  17. [17]
    Navigating the Complexities of Material Failures - Envista Forensics
    May 6, 2024 · Overloading happens when the stress on an application exceeds the material's strength, often resulting in quick fracture surfaces. Overloading ...
  18. [18]
    The exponential law of endurance tests - Semantic Scholar
    Semantic Scholar extracted view of "The exponential law of endurance tests" by O. Basquin.Missing: original | Show results with:original
  19. [19]
    A complete guide to Metal Failure Modes and Analysis - BES Group
    Feb 26, 2025 · Creep failure is the gradual deformation of metals over a long period of time when it is exposed to constant stress at high temperatures. Under ...
  20. [20]
    Strength and Stiffness Characteristics
    The maximum tensile stress a material is capable of carrying is known as the ultimate tensile stress. It corresponds to the highest point on the tensile stress ...
  21. [21]
    [PDF] Design and Evaluation of Steel Bridges for Fatigue and Fracture
    Examples of failures were reported in elements such as ... Isolated failures of riveted truss members due to fatigue have been reported in the literature.
  22. [22]
    [PDF] 1 CHAPTER 5 CORROSION TYPES 5.1 Uniform ... - Naval Academy
    5.2 Galvanic Corrosion. When two different metals are exposed to a corrosive environment, an electrical potential difference will exist.Missing: Nernst equation
  23. [23]
    21 Types of Pipe Corrosion & Failure - Corrosionpedia
    Nov 9, 2020 · Failures can be minor (in the form of a pinhole leak) or catastrophic, with significant losses due to water damage as well as the cost of pipe ...
  24. [24]
    [PDF] AC 43-4B - Advisory Circular
    Nov 9, 2018 · For example, aluminum skin panels and stainless steel doublers, riveted together in an aircraft wing, form a galvanic couple if moisture and ...<|control11|><|separator|>
  25. [25]
    [PDF] Pipeline Corrosion
    Risers will fail often, but the failure is rarely catastrophic and downtime is usually minimal as compared with onshore pipeline failures due to corrosion.
  26. [26]
    5 Most Common Factors Affecting Corrosion - Corrosionpedia
    Mar 14, 2023 · When metals are exposed to factors such as diffusion, conductivity, pH imbalance, electrochemical potential and certain types of ions, corrosion may take place ...
  27. [27]
    Understanding the Nernst Equation - Gamry Instruments
    The Nernst Equation is a cornerstone of electrochemical analysis. It quantifies how the potential of an electrochemical cell shifts based on the concentration ...<|separator|>
  28. [28]
    Nerst Equation - Corrosionpedia
    The Nernst equation is a formula that represents the relationship between an electrode potential and the solution concentration.
  29. [29]
    Corrosion Science Knowledge Area | WBDG
    Oct 7, 2021 · Environmental Severity Factors Influencing Corrosion · Temperature · Time of wetness and humidity · Presence of chlorides · pH · Industrial ...
  30. [30]
  31. [31]
    Fundamentals of arc flash mitigation - Eaton
    Damaged insulation, loose or incorrect wiring, and faulty control connections can cause failures and arc flash incidents. Insulation breakdown can lead to ...
  32. [32]
    Why substation equipment fails and why it's wise to think of that ...
    Dec 9, 2019 · Typical faults are loose joints leading to overheating or arcing and the existence of voids and contamination in insulation causing arcing and ...
  33. [33]
    Basics of Thermal Dissipation - Definition, Mechanism, and Equation ...
    When temperatures rise above the thermal limit, the component burns out and, as discussed, it is one of the most common causes of electronic component failure.Missing: expansion | Show results with:expansion
  34. [34]
    Thermal Fatigue - an overview | ScienceDirect Topics
    Thermal fatigue is a fatigue failure with macroscopic cracks resulting from cyclic thermal stresses and strains due to temperature changes.Missing: dissipation | Show results with:dissipation
  35. [35]
    Joule's Law of Electric Heating | Advanced PCB Design Blog
    Apr 12, 2023 · Joule's law states heat generated is proportional to wire resistance, time, and the square of current (Q = I²Rt).
  36. [36]
    What Causes Transformer Overheating and How to Prevent It?
    Nov 5, 2024 · Transformer overheating can significantly compromise both performance and safety, with root causes ranging from overloading and insulation failures to ...
  37. [37]
    What is “Thermal Runaway” in Industrial Electronics?
    Thermal runaway is a chain reaction. As a component heats up, its electrical resistance may change. This leads to more current flow and even more heat ...
  38. [38]
    Insulation Failure - an overview | ScienceDirect Topics
    Material corrosion/degradation can occur naturally, or be accelerated through electrical stimulation. Corrosion products have concerns of toxicity [7], but ...
  39. [39]
    Cascading Failures: Reducing System Outage - Google SRE
    For example, a single replica for a service can fail due to overload, increasing load on remaining replicas and increasing their probability of failing, causing ...
  40. [40]
    8.1.2.4. "Bathtub" curve - Information Technology Laboratory
    A plot of the failure rate over time for most products yields a curve that looks like a drawing of a bathtub, If enough units from a given population are ...Missing: engineering equations
  41. [41]
    None
    ### Summary of Bathtub-Shaped Failure Rate in Mechanical Components
  42. [42]
    How To Tell If A Car Engine Is Seized | 1800 Salvage
    Continuous overheating due to a malfunctioning cooling system can cause the engine to seize. Piston ring failure. Damaged or broken piston rings can lead to ...
  43. [43]
    [PDF] Fault Tree Analysis - NASA Technical Reports Server (NTRS)
    by a gate (e.g., AND, OR). The tree consists of primary events, secondary events, and logic gates. A major. 63. Page 90. goal of fault tree analysis is to ...
  44. [44]
    How to Do Root Cause Analysis (RCA) - Learn Lean Sigma
    The 6-Step Root Cause Analysis Process · 1: Define the Problem · 2: Collect Data · 3: Identify Possible Causes · 4: Determine the Root Cause(s) · 5: Implement ...What Is Root Cause Analysis? · A Brief History Of Rca · Rca In The Dmaic Framework<|separator|>
  45. [45]
    What are the Five Whys? A Tool For Root Cause Analysis | Tulip
    The Five Whys method was originally developed by Sakichi Toyoda, the founder of Toyota Industries. This method became widely used in Toyota Motor Corporation ...
  46. [46]
    [PDF] Ishikawa Root Cause Analysis Methodology - OPM
    The Fishbone Analysis ... What is an Ishikawa Diagram? • Dr. Kaoru Ishikawa, noted for his quality management innovations, invented the fishbone diagram in 1968.
  47. [47]
  48. [48]
    ReliaSoft BlockSim: Reliability, Availability, and Maintainability ...
    BlockSim supports an extensive array of reliability block diagram (RBD) configurations and fault tree analysis (FTA) gates and events, including advanced ...
  49. [49]
    [PDF] Fail-Safe and Safe-Life Designs And Factor of Safety Factors of ...
    The purpose of using a safety factor is to assure that the design does not fail in the event. of unexpectedly high loads or the presence of material/design ...
  50. [50]
    [PDF] Interrelation Between Safety Factors and Reliability
    The safety factor can be thought of as a measure of the designer's. NASA/CR--2001-211309. 2. Page 6. uncertainty in the analytical models, failure theories, and ...
  51. [51]
  52. [52]
    All You Need to Know About Modularization
    Modularization is a tool to subdivide the product in a logical way that provides both economies of scale and the ability to create different configurations.
  53. [53]
    Using Finite Element Analysis to Assess and Prevent the Failure of ...
    Dec 7, 2016 · FEA is used to study how a structure will respond to complex stress loading and to predict any possibility of structural failure.
  54. [54]
    Derating - RMQSI Knowledge Center
    Derating is limiting stresses on parts below their ratings, creating a margin between applied stress and maximum rating to reduce failure probability.
  55. [55]
    ISO 26262-1:2018 - Road vehicles — Functional safety — Part 1
    In stockISO 26262-1:2018 defines vocabulary for safety-related E/E systems in road vehicles, providing a framework for functional safety development.
  56. [56]
    MIL-HDBK-338 Electronic Reliability Design Handbook | www.dau.edu
    The handbook is written for reliability managers and engineers and provides guidance in developing and implementing a sound reliability program for all types ...