Fact-checked by Grok 2 weeks ago

Rotational energy

Rotational is the energy possessed by an object or system due to its about an , analogous to translational but expressed in terms of angular quantities. It is given by the formula K = \frac{1}{2} I \omega^2, where I is the —a measure of the object's resistance to depending on its distribution relative to the —and \omega is the in radians per second. This form arises from integrating the translational kinetic energies of all constituent particles in a , where each particle's linear speed v_i = r_i \omega and the I = \sum m_i r_i^2, yielding the compact rotational expression. For a in general motion, the total is the sum of the translational of its and the rotational about that center, K_{\text{total}} = \frac{1}{2} M v_{\text{cm}}^2 + \frac{1}{2} I_{\text{cm}} \omega^2, where M is the total mass and v_{\text{cm}} is the . This separation simplifies in scenarios like rolling without , where rotational and translational components are linked by v_{\text{cm}} = r \omega. The concept extends the work-energy theorem to rotations, stating that the net work done by torques equals the change in rotational , W = \int \tau \, d\theta = \Delta K, enabling predictions of motion without detailed force tracking. Rotational kinetic energy is fundamental in diverse fields, underpinning conservation laws in isolated systems where angular momentum and energy remain constant absent external torques or non-conservative forces. In , it informs the design of flywheels for in vehicles and machinery, as seen in experimental buses that recapture braking energy as rotational motion. For instance, a (four blades, total moment of inertia ≈ 1070 kg·m²) spinning at 300 rpm stores approximately 5.3 × 10⁵ J of rotational . In , it models the dynamics of human limbs, joints, and tools during activities like swinging a or walking, where rotational motion contributes significantly to overall expenditure and . These applications highlight its role in optimizing performance, from rotors to like boomerangs thrown at 30 m/s while rotating at 10 revolutions per second, which combine rotational (≈ 490 J) and translational (450 J) energies totaling approximately 940 J.

Fundamentals

Definition

Rotational energy, commonly known as , is the portion of possessed by an object due to its about an , setting it apart from translational that results from the straight-line motion of the object's . This form of energy captures the dynamic state of rigid bodies in rotational motion, where the overall includes contributions from both and any accompanying . Physically, rotational energy manifests in the stored potential for work from spinning objects, such as wheels or , originating from the tangential velocities of individual elements orbiting the axis. In the (SI), it is quantified in joules (J), defined as 1 J = 1 kg·m²/s², consistent with the units for all forms of . The concept of rotational energy was formalized in the through advancements in , primarily by Leonhard Euler, who developed key principles between 1738 and 1775, and later refined by . It depends on factors like the , a measure of how mass is distributed relative to the axis, and , the rate of rotational change.

Comparison to Translational Kinetic Energy

Rotational kinetic energy serves as a direct counterpart to translational , representing the energy associated with an object's motion but replacing linear quantities with angular ones. Just as translational quantifies the work needed to accelerate a linearly, rotational kinetic energy does the same for around an , with the acting as the rotational analog to and to linear . A primary difference lies in how these energies depend on an object's properties: translational kinetic energy relies solely on the total and the of the center of , independent of shape or internal , whereas rotational kinetic energy hinges on the of relative to the of , making it sensitive to the object's and the chosen . For instance, two objects of equal moving at the same linear speed have identical translational , but if rotating, a more stores less rotational energy than a spread-out one due to its smaller . This contrast highlights rotation's complexity, as varies with how is arranged around the pivot point. For rigid bodies undergoing combined motion, the total is the sum of the translational component—based on the center-of-mass velocity—and the component—based on about that center. This additive nature means that pure ignores rotational contributions, while scenarios involving both, such as an object tumbling through space, account for each separately to yield the full profile. Intuitively, consider a sliding down an incline versus a rolling down the same path: the sliding converts all into translational , moving faster overall, while the rolling splits that energy between translation of its center and about it, resulting in slower linear progress but added from the rotational component. Similarly, a spinning top at rest has only rotational energy, contrasting with a thrown that gains translational energy without initial , illustrating how enhances or modifies total motion without altering the underlying principles.

Mathematical Formulation

Rotational Kinetic Energy Formula

The rotational kinetic energy of a rotating about a fixed is given by the formula E_k = \frac{1}{2} I \omega^2, where I is the of the body about the axis of , and \omega is the measured in radians per second. This expression represents the instantaneous associated with the at a given \omega. The I quantifies the body's resistance to and is defined as I = \int r^2 \, dm, where r is the from the to the mass element dm./9%3A_Rotational_Kinematics_Angular_Momentum_and_Energy/9.5%3A_Rotational_Kinetic_Energy) This formula applies under the assumptions of rigid body rotation about a fixed axis, where all parts of the body maintain constant distances from the axis and there is no deformation or translation of the center of mass. It is analogous to the translational kinetic energy formula \frac{1}{2} m v^2, with I playing the role of mass m and \omega that of linear velocity v./9%3A_Rotational_Kinematics_Angular_Momentum_and_Energy/9.5%3A_Rotational_Kinetic_Energy) For single-axis rotation, the energy is a scalar quantity. In the more general case of three-dimensional rigid body motion, the rotational kinetic energy takes a vector form involving the inertia tensor: E_k = \frac{1}{2} \boldsymbol{\omega} \cdot \mathbf{I} \cdot \boldsymbol{\omega}, where \boldsymbol{\omega} is the vector and \mathbf{I} is the tensor./04%3A_Rigid_Body_Rotation/4.03%3A_Kinetic_Energy_of_Rigid_Body_Rotation) An extension of the work-energy theorem to rotational motion states that the change in rotational equals the work done by the net torque: \Delta E_k = \int \tau \, d\theta, where \tau is the torque and \theta is the angular displacement.

Moment of Inertia

The , often denoted as I, quantifies an object's resistance to changes in its rotational motion about a specific , depending on the distribution of its relative to that . For a continuous , it is defined mathematically as the I = \int r^2 \, dm, where r is the from the of to the element dm. This formulation arises from considering the as composed of point masses, each contributing r^2 \, dm to the total, emphasizing how farther from the increases I more significantly. Physically, the moment of inertia serves as the rotational analog to mass in , determining the required to produce a given . A larger implies greater rotational kinetic for the same \omega, as expressed in the energy formula \frac{1}{2} I \omega^2. To compute the about axes other than through the center of mass, the parallel axis states that I = I_{cm} + M d^2, where I_{cm} is the about the center-of-mass axis parallel to the new axis, M is the total mass, and d is the between the axes. This facilitates calculations for arbitrary axes by relating them to the more straightforward center-of-mass . For planar objects (laminae) in the xy-plane, the provides I_z = I_x + I_y, where I_z is the about the z-axis (perpendicular to the ) passing through the same point, and I_x, I_y are about the x- and y-axes, respectively. This relation is particularly useful for two-dimensional mass distributions, allowing computation of one from the other two.

Theoretical Foundations

Derivation from First Principles

The rotational of a arises directly from the translational of its elements undergoing about a fixed . To derive this, consider a rotating with constant angular velocity \omega. Each small element dm located at a r from the of has a tangential speed v = r \omega. The contribution of this element is dK = \frac{1}{2} dm \, v^2 = \frac{1}{2} dm (r \omega)^2 = \frac{1}{2} r^2 \omega^2 \, dm. Summing over all mass elements requires integrating this expression across the body: K = \int \frac{1}{2} r^2 \omega^2 \, dm = \frac{1}{2} \omega^2 \int r^2 \, dm. The integral \int r^2 \, dm defines the I about the axis, yielding the standard formula K = \frac{1}{2} I \omega^2. This approach builds on the familiar translational \frac{1}{2} m v^2 by accounting for the distributed nature of rotation. This derivation assumes the body is rigid, with all points maintaining fixed relative positions (no deformation), and rotation occurs about a fixed , ensuring uniform for all elements. It applies within non-relativistic , where velocities are much less than the and quantum effects are negligible. An alternative perspective emerges in the Lagrangian formulation of mechanics, where the system's dynamics are described using . For pure about a fixed , the angle serves as the generalized coordinate, with angular velocity . The kinetic energy term in the L = T - V (where V is ) becomes T = \frac{1}{2} I \dot{\theta}^2, mirroring the derived expression and facilitating the use of Lagrange's equations for more complex systems.

Relation to Angular Momentum

Angular momentum \mathbf{L} for a rigid body rotating about a fixed is defined as the product of its I and \omega, given by \mathbf{L} = I \omega, where \mathbf{L} is a vector directed along the axis of rotation according to the . The rotational E_k relates directly to through the expression E_k = \frac{L^2}{2I}, which follows from substituting L = I \omega into the standard energy formula \frac{1}{2} I \omega^2; this form highlights how, for a fixed moment of inertia, the energy depends quadratically on the . In systems where the varies, such as a figure skater pulling in their arms to spin faster, is conserved if no external acts, so L = I \omega = constant implies that a decrease in I increases \omega, thereby increasing the rotational despite conservation of L. \tau governs changes in via \tau = \frac{dL}{dt}, and the power delivered to the rotating system is P = \tau \omega, which equals the rate of change of rotational \frac{dE_k}{dt}.

Practical Examples

Everyday Objects

Rotational energy manifests in everyday objects through their spinning or rolling motions, where the associated with combines with translational motion in many cases. For a solid rolling without slipping, the total kinetic energy is the sum of translational and rotational components: \frac{1}{2} m v^2 + \frac{1}{2} I \omega^2, where v is the linear speed, \omega is the angular speed related by v = r \omega, and I = \frac{2}{5} m r^2 is the for a solid about its center. A spinning provides a relatable example of pure rotational energy. Approximating the wheel's as about 0.22 kg·m² for a typical model with distributed between the and , and an speed of roughly 20 rad/s during moderate pedaling, the rotational is \frac{1}{2} I \omega^2 \approx 48 J. This energy level illustrates how even modest rotations store noticeable amounts of energy, contributing to the wheel's stability and during cycling. Flywheels harness rotational for practical , converting electrical input into high-speed to smooth fluctuations in systems like engines or grids. The stored , given by \frac{1}{2} I \omega^2, is released on demand by slowing the flywheel, providing rapid bursts of without chemical reactions. This mechanical approach is efficient for short-term energy buffering, as seen in automotive applications where flywheels maintain consistent rotation during cycles. On a larger scale, Earth's daily exemplifies immense rotational energy. With a I \approx 8.04 \times 10^{37} kg·m² and angular speed \omega = 7.29 \times 10^{-5} rad/s, the planet's rotational is approximately $2.14 \times 10^{29} J. This vast quantity underscores the scale of rotational dynamics in familiar celestial contexts.

Planetary and Astronomical Bodies

The rotational kinetic energy of Earth about its axis is given by K = \frac{1}{2} I \omega^2, where I is the moment of inertia and \omega is the angular velocity. Earth's moment of inertia is approximately $8.04 \times 10^{37} kg m², reflecting its layered density structure with a denser core. The angular velocity is \omega = 7.292 \times 10^{-5} rad/s, corresponding to one rotation every sidereal day. Substituting these values yields K \approx 2.14 \times 10^{29} J. This energy drives geophysical processes like the maintenance of the magnetic dynamo, though it is gradually dissipated through tidal interactions. Tidal friction from the Earth- system causes to slow, increasing the length of the day by about 2.3 ms per century. This deceleration arises primarily from gravitational interactions with the , which raise tidal bulges on that lag slightly behind the planet's rotation due to ; the 's gravity then exerts a on these bulges, transferring from 's spin to the 's orbit. Over billions of years, this process has lengthened the day from around 6 hours in the early solar system and will continue, potentially leading to in the distant future. Among planets, exemplifies high rotational energy due to its rapid spin and massive size. Its sidereal rotation period is 9.925 hours, yielding an of approximately $1.76 \times 10^{-4} rad/s. 's is $2.45 \times 10^{42} kg m², determined from gravitational measurements by the mission, with a normalized value of about 0.264 MR^2 accounting for its oblate shape and internal density gradient. The resulting rotational kinetic energy is roughly $3.8 \times 10^{34} J, orders of magnitude greater than Earth's, powering phenomena like its intense and atmospheric dynamics. In astrophysical contexts, rotational energy of compact objects like s can be extracted via the . Proposed in 1969, this mechanism involves a particle entering the of a rotating (Kerr) , where drags faster than light speed; the particle splits into two, with one falling into the hole carrying while the other escapes with more energy than the original, effectively extracting up to 20.7% of the 's total rotational energy (the irreducible mass limit). This process highlights how a 's spin parameter a (up to a = 1 for maximal rotation) stores immense energy, E \approx 0.29 Mc^2 for extremal cases, influencing phenomena like jet formation in active galactic nuclei. Stellar remnants such as pulsars store and release rotational energy through radiation and other mechanisms. For example, the , a young with a 33-millisecond period, loses rotational at a rate of about $4.5 \times 10^{31} erg/s, powering the surrounding Crab Nebula's optical and emissions. This energy originates from the pulsar's initial spin, with total rotational energy on the order of $10^{46} erg, gradually decreasing as the period lengthens. In accretion disks around or , rotational energy dominates the dynamics, with gas in near-Keplerian orbits converting orbital to via viscous dissipation; for a stellar-mass , this can release up to 42% of the rest mass energy as in efficient thin-disk models.

Applications and Extensions

Engineering and Mechanics

In engineering and mechanics, rotational energy plays a critical role in the design and operation of systems that require precise , efficient , and reliable . Engineers the rotational [formula, E](/page/Formula_E) = \frac{1}{2} I \omega^2, where I is the and \omega is , to compute and optimize in rotating components, while considering the to minimize unwanted vibrations and enhance stability in mechanical designs. Gyroscopes, particularly control moment gyroscopes (CMGs), utilize stored to maintain orientation without propellant expenditure. In systems like those on the , CMGs store up to 14,000 ft⋅lbf⋅s of through spinning rotors at 6600 rpm, generating via adjustments to counteract disturbances and preserve stability. and generators convert from fluid into electrical power, with wind exemplifying efficiency limits in this process. The Betz limit establishes that no can extract more than 59.3% of the wind's into on the blades, derived from assuming undisturbed downstream and an optimal ratio of one-third slowdown. In practice, modern three-bladed designs achieve 75-80% of this theoretical maximum, translating to overall system efficiencies of around 45-47% when accounting for generator losses. In , rotational energy is integral to and enhancements, particularly in wheels and engines where it dictates dynamics and energy dissipation. systems in hybrid and electric vehicles capture this energy during deceleration by reversing the to act as a , converting the wheels' rotational kinetic energy—derived from vehicle —into that recharges the , recovering up to 50% of otherwise lost braking energy and improving by 10-25% in urban driving. Safety considerations in rotational energy systems focus on mitigating hazards from sudden energy release, especially in flywheel-based storage where high-speed rotors (7,500-50,000 rpm) store substantial . Rotor bursts or disintegration can propel fragments at lethal velocities, as seen in incidents like the 1995 testing failure causing one fatality and the 2015 explosion injuring five while damaging infrastructure. Mitigation strategies include designing with a 2.0 margin via qualification tests at 1.43 times operational speed, employing composite liners to absorb burst , and using compressive pre-stress in rotors to prevent tensile failures, ensuring controlled energy dissipation.

Advanced Contexts

In , the rotational energy of diatomic molecules is quantized, with energy levels given by E_J = \frac{[\hbar](/page/H-bar)^2}{2I} J(J+1), where J is the taking non-negative integer values, I is the , and \hbar is the . This model approximates the molecule as two point masses separated by a fixed distance, yielding discrete energy states that serve as the foundation for . probes these transitions, typically in the or far-infrared region, by measuring or spectra that reveal molecular structure, bond lengths, and isotopic compositions through the spacing between levels, which is proportional to the rotational constant B = \frac{[\hbar](/page/H-bar)^2}{2I}. For non-rigid bodies, such as vibrating diatomic molecules, vibrational-rotational introduces corrections to the levels, as centrifugal forces during rotation stretch the bond, increasing I and thus reducing the effective rotational at higher J. This manifests in spectra as deviations from equal spacing, with terms like centrifugal D modifying the to E_J \approx B J(J+1) - D [J(J+1)]^2, allowing extraction of vibrational frequencies and from rovibrational bands. In relativistic regimes, high rotational speeds near the require corrections to the classical expressions for I and \omega, arising from Lorentz contraction and frame-dependent mass increases, which alter the distribution in rigid bodies. For instance, in , the rotational energy includes higher-order terms from the , such as and mass-velocity corrections, that shift vibration-rotation levels in heavy molecules by fractions of a . In , the describes around rotating black holes, where rotational energy contributes up to 29% of the total mass-energy, extractable via processes like the Penrose mechanism, with the angular momentum parameter a parameterizing the . At atomic and molecular scales, classical rotational energy descriptions fail when quantum effects dominate, as the de Broglie of rotating particles becomes comparable to the system's size, leading to zero-point rotational energy and tunneling between states that preclude continuous \omega. The classical formula \frac{1}{2} I \omega^2 approximates the quantum levels only in the high-J limit, where semiclassical behavior emerges.

References

  1. [1]
    Rotational Kinetic Energy - HyperPhysics
    The kinetic energy of a rotating object is analogous to linear kinetic energy and can be expressed in terms of the moment of inertia and angular velocity.
  2. [2]
    10.4 Moment of Inertia and Rotational Kinetic Energy
    In this section, we define two new quantities that are helpful for analyzing properties of rotating objects: moment of inertia and rotational kinetic energy.
  3. [3]
    71 Rotational Kinetic Energy: Work and Energy Revisited
    Rotational kinetic energy has many applications. In biomechanical systems, it describes the motion of limbs, tools, and even the internal workings of joints. In ...
  4. [4]
    70. 10.4 Rotational Kinetic Energy: Work and Energy Revisited
    The only difference between rotational and translational kinetic energy is that translational is straight line motion while rotational is not. An example of ...
  5. [5]
    SP 330 - Section 2 - National Institute of Standards and Technology
    Aug 21, 2019 · It is defined by taking the fixed numerical value of the Planck constant h to be 6.626 070 15 × 10−34 when expressed in the unit J s, which is ...
  6. [6]
    [PDF] Euler and the dynamics of rigid bodies
    Note that this shows that is indeed the rotation kinetic energy of the solid. The kinetic energy with respect to an observer for which is, according to the ...
  7. [7]
  8. [8]
    Rotational Kinetic Energy – AP Physics 1 Study Guide
    Sep 17, 2024 · Rotational kinetic energy is not a new type of energy, it is a type of kinetic energy, analogous to translational kinetic energy · On a rotating ...<|separator|>
  9. [9]
    Rotational energy
    The kinetic energy of an object with translational and rotational motion is the sum of its translational and its rotational kinetic energy. Translational ...
  10. [10]
    Combined translational and rotational motion - Richard Fitzpatrick
    ... rotational kinetic energy, whereas, in the latter case, all of the released potential energy is converted into translational kinetic energy. Note that, in ...
  11. [11]
    10.4 Moment of Inertia and Rotational Kinetic Energy - OpenStax
    Sep 19, 2016 · K = 1 2 I ω 2 . 10.18. We see from this equation that the kinetic energy of a rotating rigid body is directly ...Missing: authoritative source
  12. [12]
    10.5 Calculating Moments of Inertia – University Physics Volume 1
    In integral form the moment of inertia is I = ∫ r 2 d m . Moment of inertia is larger when an object's mass is farther from the axis of rotation. It is possible ...
  13. [13]
    Center of Mass; Moment of Inertia - Feynman Lectures - Caltech
    So it is a good guess that the total moment of inertia about any axis will be I=Ic+MR2CM. This theorem is called the parallel-axis theorem, and may be easily ...
  14. [14]
    [PDF] Lagrangian Mechanics
    It is often helpful to first write the kinetic energy in Cartesian coordinates for each particle before converting to generalized coordinates.
  15. [15]
    11.2 Angular Momentum – University Physics Volume 1
    The angular momentum l → = r → × p → of a single particle about a designated origin is the vector product of the position vector in the given coordinate system ...
  16. [16]
    Energy and angular momentum
    ... rotational kinetic energy is K = ½Iω2. Rotational kinetic energy = ½ moment of inertia * (angular speed)2. When the angular velocity of a spinning wheel ...
  17. [17]
    A Race: Rolling Down a Ramp - Physics
    The final kinetic energy is made up of translational and rotational kinetic energies. Mgh = ½ Mv2 + ½ Iω2. Plugging in I = cMR2: Mgh = ½ Mv ...
  18. [18]
    Flywheels - Kinetic Energy - The Engineering ToolBox
    The kinetic energy of the rotating bicycle wheel can then be calculated to. Ef = 0.5 (0.22 kg m2) (20.9 rad/s)2. = 47.9 J. My Short List. Delete! ADD the ...
  19. [19]
    [PDF] Lab 9 - Moment of Inertia - Rotational Energy
    A sample bicycle wheel has mass 2.0 kg, half of which is in the central cylinder. The other half is in the rim and tire. The central cylinder has a radius of ...
  20. [20]
    Flywheel Energy Storage - an overview | ScienceDirect Topics
    Flywheel energy storage (FES). Flywheel energy storage or FES is a storage device which stores/maintains kinetic energy through a rotor/flywheel rotation.
  21. [21]
    9.5: Rotational Kinetic Energy - Physics LibreTexts
    Jan 14, 2019 · The Earth has a moment of inertia, I = 8.04×1037 kg·m2. Therefore, it has a rotational kinetic energy of 2.138×1029 J. The Rotating Earth: The ...
  22. [22]
    3.2: Layered Structure of a Planet - Geosciences LibreTexts
    Nov 5, 2024 · The bulk density of the earth is about \(\rho_{bulk}=5550\) kg/m\(^3\) and the average moment of inertia is \(0.33 MR_e^2\). These two models ...Moment of Inertia · Example: Trade-off between... · Moment of Inertia of Ellipsoidal...
  23. [23]
    Angular Speed of the Earth - The Physics Factbook - hypertextbook
    The Earth rotates at a moderate angular velocity of 7.2921159 × 10−5 radians/second. Venus has the slowest angular velocity (2.99 × 10−7rad/sec) and Jupiter has ...
  24. [24]
  25. [25]
    Measurement of the Earth's rotation: 720 BC to AD 2015 - Journals
    Dec 1, 2016 · This is significantly less than the rate predicted on the basis of tidal friction, which is +2.3 ms per century. Besides this linear change ...
  26. [26]
    Earth-Moon Interaction
    So if the Moon gains angular momentum, the Earth must lose angular momentum. The Earth's rotation rate is gradually slowing!
  27. [27]
    Gravity, Moment of Inertia, Core Properties in Jupiter
    Mar 24, 2021 · Here we present empirical structural models of Jupiter where the density profile is constructed by piecewise-polytropic equations.<|separator|>
  28. [28]
    Extracting black-hole rotational energy: The generalized Penrose ...
    Oct 28, 2013 · The necessary and sufficient condition for the Penrose process to extract energy from a rotating black hole is absorption of particles with negative energies ...
  29. [29]
    The thermodynamics of black holes: from Penrose process to ... - NIH
    Aug 5, 2021 · In 1969, Roger Penrose proposed a mechanism to extract rotational energy from a Kerr black hole. With this, he inspired two lines of ...The Penrose Process · Fig. 1 · Bekenstein And The Three...
  30. [30]
    Chapter 6 Pulsars
    The absorbed radiation deposits energy in the surrounding nebula, the Crab Nebula (Figure 8.10) being a prime example.6.1 Pulsar Properties · 6.1. 2 Neutron Star Masses... · 6.2 Pulsars And The...
  31. [31]
    Introduction to Accretion Disks - The Astrophysics Spectator
    Accretion disks exist because gas falling onto a gravitating object inevitably has some angular momentum that forces it into orbit around the object. Once in ...
  32. [32]
    [PDF] International Space Station Attitude Control and Energy Storage ...
    Spinning cylinders, such as reaction wheels and control moment gyroscopes, are often employed to control orientation without expending propellant. Although.
  33. [33]
    The Betz Limit for Wind Turbine Power - Alternative Energy Tutorials
    Alternative Energy Tutorial about the Betz Limit which defines the theoretical maximum efficiency of 59.3% from an open wind turbine generator.
  34. [34]
    How Regenerative Braking Works - Auto | HowStuffWorks
    A kind of braking system that can recapture much of the car's kinetic energy and convert it into electricity, so that it can be used to recharge the car's ...
  35. [35]
    [PDF] Recommended Practices for the Safe Design and Operation of ...
    Flywheels used as power management systems are not unlike a wide variety of machines operating with the potential for hazard. Excellent, comprehensive safety ...
  36. [36]
    [PDF] Challenges and Solutions for the Use of Flywheel Energy Storage in ...
    A critical step in flywheel safety is the use of prudent ... remaining rotational energy in a controlled fashion to limit loading on the flywheel housing and ...<|separator|>
  37. [37]
    [PDF] CHM 504 - Rotational Spectroscopy - The Weichman Lab
    E(J) = BJ(J + 1) − DJ J2(J + 1)2. (79). This accounts for the fact that as a real molecule rotates the bond stretches out, increasing the moment of inertia ...Missing: \hbar^ | Show results with:\hbar^
  38. [38]
    Rotational Spectroscopy of Diatomic Molecules
    Spectra of rotational transitions for diatomic and triatomic molecules measured using widely-tunable monochromatic THz source. p. 1. CrossRef · Google Scholar ...Missing: primary | Show results with:primary
  39. [39]
    Ab initio effective rotational and rovibrational Hamiltonians for non ...
    We present a perturbative method for ab initio calculations of rotational and rovibrational effective Hamiltonians of both rigid and non-rigid molecules.
  40. [40]
    Higher-order relativistic corrections to the vibration–rotation levels of ...
    The corrections considered include the two-electron Darwin, the Gaunt and Breit corrections, and the one-electron Lamb shift.Missing: speeds | Show results with:speeds