Fact-checked by Grok 2 weeks ago

Torque

Torque is a physical quantity that represents the rotational equivalent of linear force, measuring the tendency of a force to cause an object to rotate about a fixed axis. It is calculated as the product of the force applied and the perpendicular distance from the axis of rotation to the line of action of the force, known as the lever arm or moment arm. In vector form, torque \vec{\tau} is given by the cross product \vec{\tau} = \vec{r} \times \vec{F}, where \vec{r} is the position vector from the axis to the point of force application and \vec{F} is the force vector, with the magnitude \tau = r F \sin \theta (where \theta is the angle between \vec{r} and \vec{F}). The direction of torque is perpendicular to both the and vectors, determined by the , and it can produce in rigid bodies according to Newton's second law for (\vec{\tau} = I \vec{\alpha}, where I is the and \vec{\alpha} is ). Torque is a vector quantity, and its typically assigns positive values to counterclockwise rotations and negative to when viewed from a standard . In , the net torque on an object is zero, meaning the sum of all torques equals the sum of all counterclockwise torques. The SI unit of torque is the -meter (N·m), equivalent to a force of one applied at a of one meter from the , though it is distinct from work or units despite sharing the same dimensions. Torque plays a fundamental role in mechanics, appearing in applications from simple levers and pulleys to complex systems like engines, where it quantifies the twisting effectiveness of forces in generating rotational motion.

Definition and Formulation

Vector Definition

In physics, torque is defined as a \vec{\tau} given by the \vec{\tau} = \vec{r} \times \vec{F}, where \vec{r} is the from a chosen point to the point of application of the \vec{F}. This formulation captures the tendency of the to produce about the , distinguishing torque from linear , which induces translational motion along the direction. The resulting torque is to the formed by \vec{r} and \vec{F}, with its magnitude proportional to the sine of the angle between them, emphasizing the rotational component of the . The direction of the torque vector follows the for cross products: extend the fingers of the right hand along \vec{r}, curl them toward \vec{F}, and the thumb indicates the direction of \vec{\tau}, which defines the axis of rotation. This convention ensures consistency in , where torque acts as a that changes sign under improper rotations like reflections. The concept of torque, historically termed the "moment of force," originated in the early with Siméon Poisson's Traité de mécanique (1811), where it was formalized in . This development built upon ancient principles, such as ' 3rd-century BCE law of the , which equated the effects of forces at different distances from a to balance rotational tendencies. A practical example is the torque exerted when pushing on a : if \vec{r} points from the to the point of push and \vec{F} is applied at an , the torque \vec{\tau} = \vec{r} \times \vec{F} tends to rotate the door around the hinge, with maximum effect when \vec{F} is to \vec{r}.

Scalar Magnitude and Direction

The scalar magnitude of the torque, denoted as |\tau|, quantifies the rotational effect of a force about a pivot point and is calculated using the formula |\tau| = r F \sin \theta, where r is the distance from the pivot to the point of force application, F is the magnitude of the force, and \theta is the angle between the position vector and the force vector. This expression derives from the magnitude of the vector cross product, capturing the component of the force that contributes to rotation. An equivalent formulation expresses the as |\tau| = F d, where d represents the moment arm, defined as the from the to the of . Geometrically, torque embodies the "twisting strength" of , which reaches its maximum value when \theta = 90^\circ, meaning acts to the lever arm, as this alignment fully leverages the distance r for . In two-dimensional scenarios, the direction of torque is specified relative to the of motion, with the torque pointing to that ; a common assigns positive values to counterclockwise rotations and negative values to ones. For instance, when using a of length r = 0.25 m to turn a , applying a force F = 40 N to the handle (\theta = 90^\circ) produces a torque magnitude of |\tau| = 0.25 \times 40 \times \sin 90^\circ = 10 N·m, tending to rotate the counterclockwise if viewed from above. If the same force is applied at an angle \theta = 45^\circ, the magnitude reduces to |\tau| = 0.25 \times 40 \times \sin 45^\circ \approx 7.07 N·m, illustrating how the angle diminishes the effective twisting.

Relations to Kinematics and Dynamics

Connection to Angular Momentum

In rotational dynamics, torque \vec{\tau} is fundamentally defined as the time rate of change of \vec{L} of a system, expressed by the equation \vec{\tau} = \frac{d\vec{L}}{dt}. This relation holds for any system of particles or rigid bodies, where the net external torque equals the total rate of change of the system's . For a rotating about a fixed , the is given by \vec{L} = I \vec{\omega}, where I is the about that and \vec{\omega} is the . Substituting this into the fundamental equation yields \vec{\tau} = I \frac{d\vec{\omega}}{dt} = I \vec{\alpha} for systems of constant mass and , where \vec{\alpha} is the . This form is the rotational analog of Newton's second law for , \vec{F} = \frac{d\vec{p}}{dt}, highlighting 's role in altering rotational motion much like changes . When no external torque acts on a system (\vec{\tau} = 0), the angular momentum remains conserved, \vec{L} = constant, leading to applications in phenomena like the steady spin of a gyroscope or the orbital motion of planets. For instance, applying torque to a wheel, such as through friction from the ground, increases its angular speed (\alpha > 0) during acceleration or decreases it during braking, directly altering its angular momentum. In variable mass systems, such as rockets expelling propellant, the relationship between torque and angular momentum becomes more involved due to the changing mass distribution, requiring additional terms to account for the momentum carried away by ejected material.

Derivatives and Time Evolution

In non-steady rotational systems, the time of torque, denoted \dot{\tau} = \frac{d\tau}{dt}, describes the instantaneous rate of change of the applied rotational , playing a key role in analyzing behaviors like oscillations or rapid accelerations where torque does not remain constant. This is particularly relevant in transient conditions, where it influences the smoothness of motion and system stability. For instance, in oscillatory systems, \dot{\tau} captures how quickly the torque adjusts to varying inertial demands or external perturbations. The relation between \dot{\tau} and higher-order kinematics derives from the core equation of rotational dynamics, \vec{\tau} = I \vec{\alpha}, where I is the moment of inertia and \vec{\alpha} is the angular acceleration. Differentiating with respect to time yields \dot{\vec{\tau}} = \dot{I} \vec{\alpha} + I \dot{\vec{\alpha}} for cases of variable inertia, with the second term involving angular jerk \dot{\vec{\alpha}} = \frac{d\vec{\alpha}}{dt}, the rate of change of angular acceleration. When I is constant—as in most rigid-body analyses—this simplifies to \dot{\vec{\tau}} = I \dot{\vec{\alpha}}, linking the torque derivative directly to jerk and highlighting its role in quantifying rotational "jolt." This formulation extends the analogy to linear dynamics, where the force derivative relates to linear jerk, and is applied in control systems to bound motion smoothness. Applications of torque derivatives appear prominently in transient dynamics. In electric motor startups, the torque evolves rapidly from an initial high value—often 150–% of rated torque—to accelerate the rotor, with \dot{\tau} governing the transition rate and preventing excessive mechanical stress; for induction motors, this profile typically peaks within milliseconds before decaying as speed approaches nominal values. Similarly, in damped pendulums, the net torque varies temporally due to gravitational restoration and viscous damping, resulting in \dot{\tau} that modulates the of oscillations, ensuring controlled . A practical example is a swinging mechanism, where applied torque must adjust dynamically to changing ; as the accelerates from rest, initial high torque overcomes static and , but \dot{\tau} becomes negative during deceleration phases influenced by hinges or air , yielding a nonlinear time profile that ensures smooth closure without slamming. In engineering design, such torque-time profiles are routinely plotted or tabulated to evaluate system performance—for instance, in automotive engines, where transient torque ramps during are optimized to balance power delivery and component durability, often showing initial surges followed by steady-state plateaus over 0.1–1 second timescales.

Energy and Power Aspects

Relation to Mechanical Power

In rotational systems, the instantaneous mechanical power P is the rate at which torque transfers energy to or from the system, given by the scalar product P = \tau \omega when the torque \tau and angular velocity \omega are aligned along the axis of rotation. More generally, for arbitrary orientations, the power is the vector dot product P = \vec{\tau} \cdot \vec{\omega}, capturing only the component of torque parallel to the angular velocity. This formulation highlights power as the instantaneous energy transfer rate driven by the torque acting at a specific angular speed. The units of mechanical power align consistently with those of torque and angular velocity: power is measured in watts (W), or joules per second (J/s), derived from torque in newton-meters (N·m) multiplied by angular velocity in radians per second (rad/s), yielding \mathrm{N \cdot m \cdot rad/s = J/s} since the radian is dimensionless. A practical example occurs in electric motors, where output power is computed from measured torque and rotational speed in revolutions per minute (RPM) using P = \tau \cdot \omega, with \omega converted via \omega = 2\pi \cdot \mathrm{RPM}/60; for instance, a motor delivering 10 N·m at 1500 RPM produces approximately 1570 W. When angular velocity varies over time, such as during or load changes, the instantaneous P(t) = \vec{\tau}(t) \cdot \vec{\omega}(t) fluctuates accordingly, whereas is the time-averaged value over an interval, providing a measure of net delivery.

Relation to Work and Energy

In rotational , the work done by a torque \vec{\tau} on a over an angular displacement \Delta \vec{\theta} is given by the W = \int \vec{\tau} \cdot d\vec{\theta}, which is analogous to the work done by in , W = \int \vec{F} \cdot d\vec{x}. This formulation arises because torque causes rotational displacement, transferring to the . For a constant torque, the expression simplifies to W = \tau \Delta \theta, where \Delta \theta is the total angular displacement magnitude. The rotational work-energy theorem states that the net work done by all torques equals the change in rotational of the body: \Delta K = \frac{1}{2} I (\omega_f^2 - \omega_i^2) = \int \tau \, d\theta, where I is the , \omega_f and \omega_i are the final and initial angular velocities, respectively. This connection highlights how torque accelerates , converting work into , much like does in translation. To derive the integral form, consider the instantaneous P = \frac{dW}{dt} = \tau \omega, combined with \omega = \frac{d\theta}{dt}, yielding dW = \tau \, d\theta; integrating over the displacement gives W = \int \tau \, d\theta. Torque can also relate to changes in potential energy, particularly in systems involving gravitational or elastic restoring torques. For instance, in a simple , the gravitational torque \tau = -mg l \sin \theta (where m is , g is , l is to the center of mass, and \theta is the from vertical) opposes motion, and the work done against this torque increases the gravitational U = m g l (1 - \cos \theta). Similarly, in a torsional system, applying torque to wind the spring stores elastic ; the work W = \int_0^\theta \kappa \theta' \, d\theta' = \frac{1}{2} \kappa \theta^2 (with \kappa as the torsional constant) directly equals this potential increase. A practical example is lifting a load using a pulley system driven by torque. Here, the applied torque \tau = r F (with r as radius and F as ) rotates the pulley through \Delta \theta, performing work W = \tau \Delta \theta = F (r \Delta \theta) = F \Delta h, where \Delta h is the linear displacement of the load; this equals the gain in gravitational potential energy m g \Delta h of the lifted m.

Units and Measurement

SI Units

In the (SI), torque is quantified using the (N·m), defined as the torque resulting from a of one applied perpendicularly over a lever arm of one . This unit physically interprets torque as the product of and , emphasizing its role in inducing about an . The shares dimensional equivalence with the (J), the SI unit of , as both are ·²·⁻² in base units; however, torque is distinctly not an measure due to its vectorial nature and dependence on the of application rather than path-integrated work. The Bureau International des Poids et Mesures (BIPM) maintains the coherence of the N·m through fixed definitions of the , , and second, ensuring and precision in metrological standards. Torque measurement employs specialized instruments, including torque wrenches for applying and verifying controlled values in tasks and dynamometers for assessing rotational forces in engines and machinery, with calibrations adhering to ISO 6789 for accuracy within ±4% of reading. In practical contexts, torque values typically span from 10^{-3} N·m in micro-mechanical devices to 10^{6} N·m in large-scale industrial systems, reflecting the unit's versatility across scales.

Conversions to Other Units

Torque, as measured in the SI unit of , can be converted to various non-SI units commonly used in and . One prevalent unit is the pound-force foot (lbf·ft), where 1 N·m = 0.737562 lbf·ft. This factor arises from the definitions of the base units: the equals 1 ·m/s², while the pound-force (lbf) equals 4.4482216152605 N exactly, and the foot (ft) equals 0.3048 m exactly. Thus, 1 lbf·ft = 4.4482216152605 N × 0.3048 m = 1.3558179483314004 N·m, so the inverse conversion yields 1 N·m ÷ 1.3558179483314004 ≈ 0.737562 lbf·ft. Another common imperial unit for torque, particularly in smaller-scale applications like fasteners, is the inch-pound force (in·lbf), with 1 N·m = 8.85075 in·lbf. This derives similarly from the inch equaling 0.0254 m exactly, making 1 in·lbf = (4.4482216152605 N × 0.0254 m) / 12 ≈ 0.112984829 N·m, and thus 1 N·m ≈ 8.85075 in·lbf. For even finer measurements, such as in precision instruments, the ounce-force inch (ozf·in) is used, where 1 ozf·in ≈ 0.007061552 N·m, so 1 N·m ≈ 141.612 ozf·in. In , torque specifications for engines are frequently expressed in lbf·ft in the United States or N·m internationally. This duality reflects regional standards, with conversions essential for global comparisons of vehicle performance data. Historical units like the kilogram-force meter (kgf·m), where 1 kgf·m = 9.80665 N·m and thus 1 N·m ≈ 0.101972 kgf·m, persist in some legacy engineering contexts but are avoided in modern scientific practice as they are non- and depend on (g ≈ 9.80665 m/s²). The SI promotes consistent use of N·m to eliminate such variability.
From N·mTo lbf·ftTo in·lbfTo ozf·inTo kgf·m
10.7375628.85075141.6120.101972
These factors are exact where derived from definitions and rounded appropriately for practical use.

Statics Principles

Principle of Moments

The states that for a in static , the sum of all torques about any chosen point must be zero, expressed as \sum \tau = 0, where \tau represents the individual torques. This ensures no net rotational tendency occurs, and it holds regardless of the pivot point selected, provided the vector sum of forces is also zero to prevent translation. This principle applies to the vector sum of torques in three dimensions but is often simplified to scalar sums for coplanar force systems, where torques are considered clockwise or counterclockwise about an axis perpendicular to the plane. The concept was formalized in by French mathematician Pierre Varignon in 1687, who provided a rigorous geometric proof for the of moments in force . Varignon's theorem states that the total torque produced by a of forces about any point equals the torque due to their about the same point. This allows simplification of complex force distributions into an equivalent single force for moment calculations, aiding analysis in static . A practical example is a or beam , where two masses m_1 and m_2 are placed at distances d_1 and d_2 from the . For , the clockwise torque equals the counterclockwise torque: m_1 g d_1 = m_2 g d_2, allowing calculation of the position d_1 = \frac{m_2}{m_1 + m_2} L for a beam of length L. This illustrates how the principle ensures in everyday systems.

Static Equilibrium Conditions

For a rigid body to be in static equilibrium, two fundamental conditions must be satisfied: the vector sum of all external forces must be zero (\sum \vec{F} = 0), which prevents translational acceleration, and the vector sum of all external torques about any arbitrary point must be zero (\sum \vec{\tau} = 0), which prevents rotational acceleration. These conditions apply to systems at rest relative to an inertial frame, encompassing both the principle of moments for torque and the balance of forces. In two-dimensional planar problems, these translate to three independent equations: two for forces (\sum F_x = 0 and \sum F_y = 0) and one for torque to the (\sum \tau_z = 0). In three dimensions, the conditions expand to six independent equations: three for forces (\sum F_x = 0, \sum F_y = 0, \sum F_z = 0) and three for torques about each (\sum \tau_x = 0, \sum \tau_y = 0, \sum \tau_z = 0). Calculating torques about two strategically chosen points in can provide the necessary independent equations to resolve components, as the full set ensures about any point when forces balance. When the number of unknown reaction forces or supports exceeds the available equilibrium equations, the system becomes , requiring supplementary relations from deformation compatibility or constitutive laws to solve. For instance, a with three supports in provides only three equations but four unknowns, necessitating additional analysis beyond pure . A classic example is a uniform leaning against a frictionless wall, where torque equilibrium about the base resolves unknown forces. Consider a 5.0-m of 10 kg at a 60° to the ground, with a 70-kg 4.0 m up the ; the torque due to the ladder's weight acts at its center (2.5 m along the ladder), the person's weight at 4.0 m, and the wall's horizontal provides a counter-torque at the top (5.0 m). Setting \sum \tau = 0 about the base eliminates and ground unknowns, yielding the wall force as approximately 345 N (using g = 9.8 m/s²), after which force balance solves the rest. Torque balance in static equilibrium ensures no net rotation, thereby preventing tipping under applied loads, but it does not guarantee resistance to small oscillations, which requires damping mechanisms in the system to restore position after perturbations. Stable equilibrium occurs when displacements produce restoring torques opposite to the deviation, as in a system with a low center of gravity.

Dynamic and Applied Cases

Moment Arm Formula

The moment arm formula provides a practical method for calculating torque in systems where a force acts at a distance from a pivot point, particularly in lever-like configurations. The magnitude of the torque \tau is expressed as \tau = F d, where F is the magnitude of the force and d is the moment arm, defined as the shortest perpendicular distance from the pivot to the line of action of the force. This formula derives from the vector definition of torque, where the is |\tau| = r F \sin \theta, with r being the from the to the point of application and \theta the angle between the position vector \mathbf{r} and the vector \mathbf{F}. The moment arm d corresponds to the component of r to \mathbf{F}, such that d = r \sin \theta, leading directly to \tau = F d. Geometrically, d is determined by constructing the from the to the extended of , akin to dropping a plumb line to find the horizontal offset or projecting onto the to the position . This construction ensures d captures the effective length contributing to , independent of 's point of application along its line. For instance, in a , a driver applies tangent to the rim, where d equals the wheel's , maximizing torque for a given F and enabling efficient turning with minimal effort; similarly, using a as a optimizes d between the and load to pry objects with reduced . The formula assumes idealized point forces acting along a straight line; for distributed loads, such as varying pressures over a surface, the total torque requires of dF \cdot d across the to account for varying arms.

Net Force Versus Torque

In , the acting on a body governs its translational motion by accelerating its according to Newton's second law for translation: \vec{a}_{\rm cm} = \frac{\vec{F}_{\rm net}}{m}, where \vec{a}_{\rm cm} is the of the , \vec{F}_{\rm net} is the vector sum of all external forces, and m is the total of the body. This equation describes how unbalanced forces cause the entire body to move linearly as if all mass were concentrated at the . In parallel, the net torque governs the body's rotational motion, producing angular acceleration about the center of mass via the rotational analog of Newton's second law: \vec{\tau}_{\rm net} = I \vec{\alpha}, where \vec{\tau}_{\rm net} is the net external torque, I is the moment of inertia about the center of mass, and \vec{\alpha} is the angular acceleration. Torque about the center of mass is the appropriate reference for this equation, as torques calculated about arbitrary points generally do not yield the same rotational dynamics unless the point coincides with the center of mass or a fixed axis through it. A single external can simultaneously produce both translational and rotational effects if applied off-center relative to the , as it contributes to \vec{F}_{\rm net} for linear and to \vec{\tau}_{\rm net} for . For instance, pushing a box horizontally through its results only in translation without , whereas an eccentric push—such as at one edge—causes the to accelerate linearly while the body also rotates about that point. This dual effect highlights how application location determines the resultant motion. A frequent misconception arises in assuming that zero net force precludes all acceleration; however, if \vec{\tau}_{\rm net} \neq 0 about of mass, the body experiences and rotates even as its center of mass remains or moves uniformly. In static , both \vec{F}_{\rm net} = 0 and \vec{\tau}_{\rm net} = 0 must hold to prevent any translation or rotation.

Torque in Machines

In and electrical machines, torque represents the rotational force generated by engines and , which is essential for loads such as wheels in vehicles or industrial equipment. Internal engines produce torque through the process, where expanding gases apply to pistons connected to a , converting into rotation. This torque varies with engine speed, forming a characteristic curve that peaks at relatively low revolutions per minute (RPM) to provide strong initial in applications like automobiles. The torque curve of a typical in a rises from low RPM, reaches a maximum around 2,000 to 4,000 RPM—where and are optimized—and then declines at higher speeds due to factors like limitations and . For electric motors, torque production arises from the between magnetic fields and in the windings. In DC motors, torque is proportional to the armature current: \tau = K_t I_a, where K_t is the torque constant depending on design parameters. For AC motors like synchronous types, torque is proportional to the sine of the load angle between stator and fields. This allows torque to be controlled by adjusting and field alignment for applications requiring precise speed-torque characteristics, such as in synchronous motors. Torque transmission in machines often involves and to adapt the output for specific needs, where the output torque \tau_\text{out} is given by \tau_\text{out} = \tau_\text{in} \times r \times \mu, with r as the (number of teeth on driven gear over driving gear) and \mu as the (typically 0.9 to 0.98 for well-lubricated systems). Higher gear ratios multiply torque at the expense of speed, enabling heavy loads like those in to overcome during startup. Dynamometers measure torque by absorbing rotational power and quantifying the reaction force, often using gauges or load cells on a torque connected to the shaft, allowing precise mapping of torque-speed relationships under controlled conditions. In automotive examples, a modern engine might deliver peak torque of around 300 , which relates to vehicle through Newton's second law for , \tau = I \alpha, where I is the rotational of the and wheels, and \alpha is ; higher torque thus enables quicker speed buildup from rest. Machine performance is often rated by , calculated as P = \tau \omega, linking torque to angular velocity \omega for overall output assessment.

Torque Multipliers

Torque multipliers are specialized tools designed to amplify the torque applied by an , enabling the tightening or loosening of fasteners that require high without excessive physical effort. These devices, such as cheater bars or geared torque multipliers, effectively extend the moment arm or employ to increase output torque. For instance, a cheater bar is a simple extension fitted over a handle to lengthen the lever arm, while more advanced torque multipliers use internal gear systems to achieve similar . The mechanics of torque multipliers rely on principles of or gear reduction to boost output torque relative to input. In lever-based designs like cheater bars, the amplification occurs by increasing the distance from the pivot point, following the relationship where output torque \tau_{out} equals input torque \tau_{in} multiplied by the of output to input distances (d_{out} / d_{in}). Geared torque multipliers, often using epicyclic or planetary gear trains, achieve higher ratios through multiple stages of gear interaction, where gear receives input, planet gears rotate, and the ring gear provides amplified output, typically yielding multiplication factors of 5 to 125 or more depending on the configuration. This gear-based approach is particularly effective for precise control in confined spaces. Common types of torque multipliers include hydraulic torque wrenches, which use fluid pressure to generate high torque for heavy-duty bolting in and ; gear reducers, such as manual or mechanical models that employ planetary gears for low-speed, high-torque applications in ; pneumatic versions powered by for rapid operation in lines; and electric models with controls for consistent performance in field service. These types cater to diverse needs, from portable hand tools to powered systems, with gear reducers excelling in scenarios requiring torque outputs up to 36,880 lbf·ft. Safety and precision are paramount in torque multiplier use, as improper application can lead to over-torquing, which damages fasteners or structures, or under-torquing, which compromises joint integrity in fastening. Devices must be calibrated regularly—typically annually or more frequently for heavy use—to maintain accuracy, often traceable to standards like those from NIST, ensuring the amplification factor remains reliable and preventing operator injury from excessive force. Quality torque multipliers reduce risks associated with makeshift tools by providing controlled, smooth operation. A practical example is removing an automobile using a with a cheater bar extension. If the standard has a 12-inch (input d_{in}) and a 24-inch cheater bar is added, the effective moment doubles (d_{out} = 2 \times d_{in}), yielding an amplification factor of 2, so 50 ft·lb of input torque produces 100 ft·lb of output torque to loosen the nut without straining the . This simple calculation highlights the device's utility in automotive maintenance.

References

  1. [1]
    10.6 Torque – General Physics Using Calculus I - UCF Pressbooks
    Figure 10.31 Torque is the turning or twisting effectiveness of a force, illustrated here for door rotation on its hinges (as viewed from overhead). Torque has ...
  2. [2]
    Torque (Moment) - NASA Glenn Research Center
    A torque produces a change in angular velocity which is called an angular acceleration. The distance L used to determine the torque T is the distance from the ...
  3. [3]
    Torque and Equilibrium - HyperPhysics
    A torque is an influence which tends to change the rotational motion of an object. One way to quantify a torque is.
  4. [4]
    Torque Direction - HyperPhysics
    It is conventional to choose it in the right hand rule direction along the axis of rotation. The torque is in the direction of the angular velocity which would ...
  5. [5]
    Archimedes' Law of the Lever
    This is the statement of the Law of the Lever that Archimedes gives in Propositions 6 and 7 of Book I of his work entitled On the Equilibrium of Planes.
  6. [6]
    Forces produce torques - Physics
    The magnitude of the torque is: τ = r' F sin(90) = r' F. Using geometry, r' = r sin(θ), so the torque (once again) has a magnitude of: τ = r F sin(θ)
  7. [7]
    [PDF] Physics 101 Lab 7: Torque
    where the force is applied is r, then the magnitude of the torque is equal to τ = rF sinθ. (2) where θ is the smallest angle between the vectors ~F and r.
  8. [8]
    Torque ( "Rotational Force" )
    The perpendicular distance from the axis of rotation (or pivot) to the "line of action" of the force F is called the "moment arm" (sometime, the "lever arm").
  9. [9]
    [PDF] Chapter 12 Torque
    Explain the causes of rotational motion using torque (the rotational analog of force). • Identify the factors that influence the ability of a force to rotate a ...
  10. [10]
    28.12 -- Wrenches, screwdrivers - UCSB Physics
    Torque, τ, equals r × F. Its magnitude is rF sin θ. Here, r is either the radius of the screwdriver handle or the length of the wrench from the center of ...
  11. [11]
    11.2 Angular Momentum - University Physics Volume 1 | OpenStax
    Sep 19, 2016 · Equation 11.8 states that the rate of change of the total angular momentum of a system is equal to the net external torque acting on the ...
  12. [12]
    3.11: Torque and Rate of Change of Angular Momentum
    Aug 7, 2022 · Theorem: The rate of change of the total angular momentum of a system of particles is equal to the sum of the external torques on the system.Missing: fundamental | Show results with:fundamental
  13. [13]
    10.7 Newton's Second Law for Rotation - University Physics Volume 1
    Sep 19, 2016 · τ = I α . The torque on the particle is equal to the moment of inertia about the rotation axis times the angular acceleration. We can ...
  14. [14]
    11.3: Angular Momentum - Physics LibreTexts
    Mar 16, 2025 · The angular momentum of a single particle about a designated origin is the vector product of the position vector in the given coordinate ...Learning Objectives · Angular Momentum of a Single... · Angular Momentum of a...
  15. [15]
    [PDF] Engineering Fundamentals - PCB Piezotronics
    ... torque-time profiles so that significant data points, which correlate with the known torque-tension relationship, are recorded. In order to properly program ...
  16. [16]
    [PDF] Dynamic stall at high Reynolds numbers due to variant types of ...
    Sep 23, 2020 · The variant types of motion of the pitching maneuvers were characterized by constant angular velocity, angular acceleration and angular jerk, ...
  17. [17]
    [PDF] Get Your Motor Running: Understanding Starting Torque - Benshaw
    Machines and motor loads have widely differing starting torque requirements. Some motor applications need only 10% rated torque to accelerate to full speed ...
  18. [18]
    A damped pendulum forced with a constant torque - AIP Publishing
    Dec 1, 2005 · In this paper, we focus on a mechanical example proposed by Andronov et al.2 consisting of a damped pendulum forced by a constant torque. This ...
  19. [19]
    Optimization of Engine Torque Management Under Uncertainty for ...
    A set of time-dependent engine torque profiles are used as input in a vehicle dynamic model (see Sec. 3), and the resulting time-dependent transmission turbine ...Torque Converter And... · Driveline And Vehicle Model · Parametric Case<|control11|><|separator|>
  20. [20]
    10.8 Work and Power for Rotational Motion - UCF Pressbooks
    The power delivered to a system that is rotating about a fixed axis is the torque times the angular velocity, P = τ ω .
  21. [21]
    Velocity Kinematics and Statics – Modern Robotics
    From physics we know that velocity times force is power, and the equivalent for a robot arm is theta-dot transpose times tau is equal to the power produced ...
  22. [22]
    [PDF] DC Motors
    Mar 27, 2010 · 1.2 Power Curve. The mechanical power P delivered by the motor is the applied torque τ times its angular velocity ˙θ. Thus. P = τ. ˙ θ = τs. ˙ θ ...
  23. [23]
    Mechanics of Materials: Torsion - Boston University
    Torque is a moment that twists a structure. Unlike axial loads which produce a uniform, or average, stress over the cross section of the object.
  24. [24]
    DC Motors
    The output power in watts is about (torque) x (rpm) / 9.57. Measuring Motor Torque. Figure 6.1: Experiment to Measure Motor Torque. \begin{figure} \fbox ...
  25. [25]
    Kinetic Energy - Physics
    ... torque and an angular displacement, work will be done by the torque. In a rotational situation: W = ∫ τ dθ. For a constant torque, W = τ Δθ. Work can be ...
  26. [26]
    [PDF] 9.5 Rotational Work and Energy
    The rotational work done by a constant torque in turning an object ... Thus, the rotational version of the Work-Energy theorem is: W. R. = KE. Rf. - KE.
  27. [27]
    The Physical Pendulum
    First we compute the net gravitational torque on the system at an arbitrary (small) angle ... Obviously its potential energy is easy enough - it depends on the ...
  28. [28]
  29. [29]
    ISO 6789-2:2017 - Assembly tools for screws and nuts
    In stockISO 6789 is applicable for the step by step (static) and continuous (quasi-static) calibration of torque measurement devices, the torque of which is defined by ...Missing: dynamometers | Show results with:dynamometers
  30. [30]
    [PDF] Guide for the Use of the International System of Units (SI)
    Feb 3, 1975 · Appendix A gives the definitions of the SI base units, while Appendix B gives conversion factors for converting values of quantities expressed ...
  31. [31]
    Technology explained, simply: Torque in cars | BMW.com
    Apr 13, 2021 · Torque is expressed in pound-feet (lb-ft) or newton-meters (Nm). The interaction of torque and engine speed (rpm) determines the engine power.
  32. [32]
    Chapter 4 Moments and Static Equivalence - Engineering Statics
    For a body in equilibrium we say that forces have a tendency to produce translation or rotation since no actual acceleration or motion occurs.
  33. [33]
    Principle of Moments | CK-12 Foundation
    You will learn how moments (torques) balance for a body to remain in rotational equilibrium, and apply this principle to various physical situations.
  34. [34]
    The Principle of Moments (Edexcel IGCSE Physics): Revision Note
    Nov 30, 2024 · The principle of moments states that: If an object is balanced, the total clockwise moment about a pivot equals the total anticlockwise moment ...<|control11|><|separator|>
  35. [35]
    Principle of Moments in Physics | Definition, Formula & Examples
    Principle of Moments, also known as Varignon's Theorem, was established by the French mathematician Pierre Varignon. The principle of moments theorem states ...Missing: statics | Show results with:statics
  36. [36]
    OpenStax | Free Textbooks Online with No Catch
    **Summary of Conditions for Static Equilibrium (OpenStax University Physics Volume 1, Section 12.2):**
  37. [37]
    OpenStax | Free Textbooks Online with No Catch
    **Summary of Ladder Example (Static Equilibrium and Torque):**
  38. [38]
    Stability – Introductory Physics for the Health and Life Sciences I
    A system is said to be in stable equilibrium if, when displaced from equilibrium, it experiences a net force or torque in a direction opposite to the direction ...
  39. [39]
    [PDF] Torque - Physics 201 Lab 10 - UNCW
    The relationships between ~F, r, and ~τ are shown in Figure 1. r⊥ = r sin θ is the moment arm of the force, and can be used to calculate the magnitude of torque ...<|control11|><|separator|>
  40. [40]
    [PDF] Chapter 10
    τ = r F sin ϕ = F d. – F is the force. – ϕ is the angle between force and r vector. – d is the moment arm (or lever arm) of the force. Page 4. Copyright © 2012 ...
  41. [41]
    [PDF] Forces and Their Measurement
    The perpendicular distance from the axis of rotation to the force line of action is known as the moment arm (d) of the force. Figure 4.6 illustrates that d ...
  42. [42]
    [PPT] Chapter 8 Rotational Motion of Solid Objects
    When the applied force is not perpendicular to the crowbar, for example, the lever arm is found by drawing the perpendicular line from the fulcrum to the line ...
  43. [43]
    9.5 Simple Machines – College Physics - UCF Pressbooks
    The perpendicular lever arms of the input and output forces are li and l0. Figure 1 shows a lever type that is used as a nail puller. Crowbars, seesaws, and ...
  44. [44]
    [PDF] Discontinuous Distributions in Mechanics of Materials - Rice University
    The distributed and resultant torques are shown in Figure 9, while the shaft rotation is given Figure 10. The corresponding input and output variables are given ...
  45. [45]
    OpenStax | Free Textbooks Online with No Catch
    **Summary of Newton's Second Law for Rotation vs. Translation (OpenStax University Physics Volume 1, Section 10.5 or 10.7):**
  46. [46]
    [PDF] Lecture L21 - 2D Rigid Body Dynamics - MIT OpenCourseWare
    For motion about the center of mass, no such restriction applies and we may obtain the statement of conservation of angular momentum about the center of mass ...
  47. [47]
    18 Rotation in Two Dimensions - Feynman Lectures - Caltech
    So we have three formulas for angular momentum, just as we have three formulas for the torque: L=xpy−ypx=rptang=p⋅lever arm. Like torque, angular momentum ...
  48. [48]
    Power vs. Torque - x-engineer.org
    In this article we are going to understand how the engine torque is produced, how engine power is calculated and what is a torque and power curve.
  49. [49]
    [PDF] Synchronous Machine and Winding Models - MIT OpenCourseWare
    The phase angle δi is called the torque angle, but it is important to use some caution, as there is more than one torque angle. Now, look at the machine from ...
  50. [50]
    Gear Reducing Formulas - The Engineering ToolBox
    Gear Output Torque. Mo = Mi r μ (1). where. Mo = output torque (Nm, lbf ft). Mi = input torque (Nm, lbf ft). r = gear transmission ratio. μ = gear efficiency ...
  51. [51]
    [PDF] DC Motor Torque-Speed Curve DC Motor Characteristics ...
    Dynamometers measure the torque and the speed of a motor. From these, power is easily calculated. It is desired to make a torque-speed curve for the motor.
  52. [52]
    What is ideal torque for sedan cars? - Quora
    Aug 7, 2016 · A small car would require around 130 to 250 NM, a medium size car 180 to 320 and a large car 240 to 400 NM. Averaging these out you get 253 NM.What is a good amount of torque for a car? - QuoraWhich is better for track racing, 100 horsepower & 300 nm torque ...More results from www.quora.com
  53. [53]
    Angular Motion - Power and Torque - The Engineering ToolBox
    Torque is a rotating force, and power is how rapidly work is done. Power is calculated as P = Tω, where T is torque and ω is angular velocity.Missing: machines | Show results with:machines
  54. [54]
    What Is a Torque Multiplier and How Does It Work?
    Jul 23, 2021 · A torque multiplier is a tool that provides a mechanical advantage when loosening or tightening nuts and bolts.Missing: engineering | Show results with:engineering
  55. [55]
    Torque Multiplier - Definition & Examples - CrossCo
    A torque multiplier is a tool engineered to simplify the process of unfastening tightly secured bolts, nuts, and various fasteners that prove resistant to ...
  56. [56]
    How Torque Multipliers Work: Gearing up for Efficiency - Micro Precision Calibration
    ### Summary of Torque Multipliers from https://microprecision.com/blog/how-torque-multipliers-work-gearing-up-for-efficiency/
  57. [57]
    All about Torque Multipliers | Michelli Weighing & Measurement
    Apr 29, 2021 · The mechanical or manual torque multipliers enable operators to use higher torque in tight spaces where a larger wrench may not be an option.Missing: definition engineering
  58. [58]