User:Brews ohare/Mechanics of planar particle motion

This article describes a particle in planar motion when observed from non-inertial reference frames.[1] [2][3][4] General results presented in fictitious forces here are applied to observations of a moving particle as seen from several specific non-inertial frames, for example, a local frame (one tied to the moving particle so it appears stationary), and a co-rotating frame (one with an arbitrarily located but fixed axis and a rate of rotation that makes the particle appear to have only radial motion and zero azimuthal motion). The Lagrangian approach to fictitious forces is introduced.

Unlike real forces such as electromagnetic forces, fictitious forces do not originate from physical interactions between objects.

Analysis using fictitious forces edit

The appearance of fictitious forces normally is associated with use of a non-inertial frame of reference, and their absence with use of an inertial frame of reference. The connection between inertial frames and fictitious forces (also called inertial forces or pseudo-forces), is expressed, for example, by Arnol'd:[5]

The equations of motion in an non-inertial system differ from the equations in an inertial system by additional terms called inertial forces. This allows us to detect experimentally the non-inertial nature of a system.

— V. I. Arnol'd: Mathematical Methods of Classical Mechanics Second Edition, p. 129

A slightly different tack on the subject is provided by Iro:[6]

An additional force due to nonuniform relative motion of two reference frames is called a pseudo-force.

— H Iro in A Modern Approach to Classical Mechanics p. 180

Fictitious forces do not appear in the equations of motion in an inertial frame of reference: in an inertial frame, the motion of an object is explained by the real impressed forces. In a non-inertial frame such as a rotating frame, however, Newton's first and second laws still can be used to make accurate physical predictions provided fictitious forces are included along with the real forces. For solving problems of mechanics in non-inertial reference frames, the advice given in textbooks is:[7] [8]

Treat the fictitious forces like real forces, and pretend you are in an inertial frame.

— Louis N. Hand, Janet D. Finch Analytical Mechanics, p. 267

Moving objects and observational frames of reference edit

Next, it is observed that time varying coordinates are used in both inertial and non-inertial frames of reference, so the use of time varying coordinates should not be confounded with a change of observer, but is only a change of the observer's choice of description. Elaboration of this point and some citations on the subject follow.

Frame of reference and coordinate system edit

The term frame of reference is used often in a very broad sense, but for the present discussion its meaning is restricted to refer to an observer's state of motion, that is, to either an inertial frame of reference or a non-inertial frame of reference.

The term coordinate system is used to differentiate between different possible choices for a set of variables to describe motion, choices available to any observer, regardless of their state of motion. Examples are Cartesian coordinates, polar coordinates and (more generally) curvilinear coordinates.

Here are two quotes relating "state of motion" and "coordinate system":[9][10]

We first introduce the notion of reference frame, itself related to the idea of observer: the reference frame is, in some sense, the "Euclidean space carried by the observer". Let us give a more mathematical definition:… the reference frame is... the set of all points in the Euclidean space with the rigid body motion of the observer. The frame, denoted  , is said to move with the observer.… The spatial positions of particles are labelled relative to a frame   by establishing a coordinate system R with origin O. The corresponding set of axes, sharing the rigid body motion of the frame  , can be considered to give a physical realization of  . In a frame  , coordinates are changed from R to R' by carrying out, at each instant of time, the same coordinate transformation on the components of intrinsic objects (vectors and tensors) introduced to represent physical quantities in this frame.

— Jean Salençon, Stephen Lyle. (2001). Handbook of Continuum Mechanics: General Concepts, Thermoelasticity p. 9

In traditional developments of special and general relativity it has been customary not to distinguish between two quite distinct ideas. The first is the notion of a coordinate system, understood simply as the smooth, invertible assignment of four numbers to events in spacetime neighborhoods. The second, the frame of reference, refers to an idealized system used to assign such numbers … To avoid unnecessary restrictions, we can divorce this arrangement from metrical notions. … Of special importance for our purposes is that each frame of reference has a definite state of motion at each event of spacetime.…Within the context of special relativity and as long as we restrict ourselves to frames of reference in inertial motion, then little of importance depends on the difference between an inertial frame of reference and the inertial coordinate system it induces. This comfortable circumstance ceases immediately once we begin to consider frames of reference in nonuniform motion even within special relativity.…the notion of frame of reference has reappeared as a structure distinct from a coordinate system.

— John D. Norton: General Covariance and the Foundations of General Relativity: eight decades of dispute, Rep. Prog. Phys., 56, pp. 835-7.

Time varying coordinate systems edit

In a general coordinate system, the basis vectors for the coordinates may vary in time at fixed positions, or they may vary with position at fixed times, or both. It may be noted that coordinate systems attached to both inertial frames and non-inertial frames can have basis vectors that vary in time, space or both, for example the description of a trajectory in polar coordinates as seen from an inertial frame.[11] or as seen from a rotating frame.[12] A time-dependent description of observations does not change the frame of reference in which the observations are made and recorded.

Fictitious forces in a local coordinate system edit

 
Figure 1: Local coordinate system for planar motion on a curve. Two different positions are shown for distances s and s + ds along the curve. At each position s, unit vector un points along the outward normal to the curve and unit vector ut is tangential to the path. The radius of curvature of the path is ρ as found from the rate of rotation of the tangent to the curve with respect to arc length, and is the radius of the osculating circle at position s. The unit circle on the left shows the rotation of the unit vectors with s.

In discussion of a particle moving in a circular orbit,[13] in an inertial frame of reference one can identify the centripetal and tangential forces. It then seems to be no problem to switch hats, change perspective, and talk about the fictitious forces commonly called the centrifugal and Euler force. But what underlies this switch in vocabulary is a change of observational frame of reference from the inertial frame where we started, where centripetal and tangential forces make sense, to a rotating frame of reference where the particle appears motionless and fictitious centrifugal and Euler forces have to be brought into play. That switch is unconscious, but real.

Suppose we sit on a particle in general planar motion (not just a circular orbit). What analysis underlies a switch of hats to introduce fictitious centrifugal and Euler forces?

To explore that question, begin in an inertial frame of reference. By using a coordinate system commonly used in planar motion, the so-called local coordinate system,[14] as shown in Figure 1, it becomes easy to identify formulas for the centripetal inward force normal to the trajectory (in direction opposite to un in Figure 1), and the tangential force parallel to the trajectory (in direction ut), as shown next.

To introduce the unit vectors of the local coordinate system shown in Figure 1, one approach is to begin in Cartesian coordinates in an inertial framework and describe the local coordinates in terms of these Cartesian coordinates. In Figure 1, the arc length s is the distance the particle has traveled along its path in time t. The path r (t) with components x(t), y(t) in Cartesian coordinates is described using arc length s(t) as:[15]

 

One way to look at the use of s is to think of the path of the particle as sitting in space, like the trail left by a skywriter, independent of time. Any position on this path is described by stating its distance s from some starting point on the path. Then an incremental displacement along the path ds is described by:

 

where primes are introduced to denote derivatives with respect to s. The magnitude of this displacement is ds, showing that:[16]

     (Eq. 1)

This displacement is necessarily tangent to the curve at s, showing that the unit vector tangent to the curve is:

 

while the outward unit vector normal to the curve is

 

Orthogonality can be verified by showing the vector dot product is zero. The unit magnitude of these vectors is a consequence of Eq. 1.

As an aside, notice that the use of unit vectors that are not aligned along the Cartesian xy-axes does not mean we are no longer in an inertial frame. All it means is that we are using unit vectors that vary with s to describe the path, but still observe the motion from the inertial frame.

Using the tangent vector, the angle of the tangent to the curve, say θ, is given by:

    and    

The radius of curvature is introduced completely formally (without need for geometric interpretation) as:

 

The derivative of θ can be found from that for sin θ:

 
 

Now:

    

in which the denominator is unity according to Eq. 1. With this formula for the derivative of the sine, the radius of curvature becomes:

  

where the equivalence of the forms stems from differentiation of Eq. 1:

 

Having set up the description of any position on the path in terms of its associated value for s, and having found the properties of the path in terms of this description, motion of the particle is introduced by stating the particle position at any time t as the corresponding value s (t).

Using the above results for the path properties in terms of s, the acceleration in the inertial reference frame as described in terms of the components normal and tangential to the path of the particle can be found in terms of the function s(t) and its various time derivatives (as before, primes indicate differentiation with respect to s):

    
 
 

as can be verified by taking the dot product with the unit vectors ut(s) and un(s). This result for acceleration is the same as that for circular motion based on the radius ρ. Using this coordinate system in the inertial frame, it is easy to identify the force normal to the trajectory as the centripetal force and that parallel to the trajectory as the tangential force.

Next, we change observational frames. Sitting on the particle, we adopt a non-inertial frame where the particle is at rest (zero velocity). This frame has a continuously changing origin, which at time t is the center of curvature (the center of the osculating circle in Figure 1) of the path at time t, and whose rate of rotation is the angular rate of motion of the particle about that origin at time t. This non-inertial frame also employs unit vectors normal to the trajectory and parallel to it.

The angular velocity of this frame is the angular velocity of the particle about the center of curvature at time t. The centripetal force of the inertial frame is interpreted in the non-inertial frame where the body is at rest as a force necessary to overcome the centrifugal force. Likewise, the force causing any acceleration of speed along the path seen in the inertial frame becomes the force necessary to overcome the Euler force in the non-inertial frame where the particle is at rest. There is zero Coriolis force in the frame, because the particle has zero velocity in this frame. For a pilot in an airplane, for example, these fictitious forces are a matter of direct experience.[17] However, these fictitious forces cannot be related to a simple observational frame of reference other than the particle itself, unless it is in a particularly simple path, like a circle.

That said, from a qualitative standpoint, the path of an airplane can be approximated by an arc of a circle for a limited time, and for the limited time a particular radius of curvature applies, the centrifugal and Euler forces can be analyzed on the basis of circular motion with that radius. See article discussing turning an airplane.

Next, reference frames rotating about a fixed axis are discussed in more detail.

Fictitious forces in polar coordinates edit

Description of particle motion often is simpler in non-Cartesian coordinate systems, for example, polar coordinates. When equations of motion are expressed in terms of any curvilinear coordinate system, extra terms appear that represent how the basis vectors change as the coordinates change.

Two terminologies edit

In a purely mathematical treatment, regardless of the frame that the coordinate system is associated with (inertial or non-inertial), extra terms appear in the acceleration of an observed particle. For example, in polar coordinates the acceleration is given by (see below for details):

 

which contains not just double time derivatives of the coordinates but added terms. This example employs polar coordinates, but the added terms depend upon which coordinate system is chosen (that is, polar, elliptic, or whatever). Sometimes these coordinate-system dependent terms also are referred to as "fictitious forces", introducing a second meaning for this term. According to this terminology, fictitious forces are determined in part by the coordinate system itself, regardless of the frame it is attached to, that is, regardless of whether the coordinate system is attached to an inertial or a non-inertial frame of reference. In contrast, the fictitious forces defined in terms of the state of motion of the observer vanish in inertial frames of reference. To distinguish these two terminologies, the fictitious forces that vanish in an inertial frame of reference are called in this article the "state-of-motion" fictitious forces and those that originate in the interpretation of time derivatives in particular coordinate systems are called "coordinate" fictitious forces.

Assuming it is clear that "state of motion" and "coordinate system" are different, it follows that the dependence of centrifugal force (as in this article) upon "state of motion" and its independence from "coordinate system", which contrasts with the "coordinate" version with exactly the opposite dependencies, indicates that two different ideas are referred to by the terminology "fictitious force". The present article emphasizes one of these two ideas ("state-of-motion"), although the other also is described.

Below, polar coordinates are introduced for use in (first) an inertial frame of reference and then (second) in a rotating frame of reference. The two different uses of the term "fictitious force" are pointed out. First, however, follows a brief digression to explain further how the "coordinate" terminology for fictitious force has arisen.

Lagrangian approach edit

}

To motivate the introduction of "coordinate" inertial forces by more than a reference to "mathematical convenience", what follows is a digression to show these forces correspond to what are called by some authors "generalized" fictitious forces or "generalized inertia forces".[18][19][20][21] These forces are introduced via the Lagrangian mechanics approach to mechanics based upon describing a system by generalized coordinates usually denoted as {qk}. The only requirement on these coordinates is that they are necessary and sufficient to uniquely characterize the state of the system: they need not be (although they could be) the coordinates of the particles in the system. Instead, they could be the angles and extensions of links in a robot arm, for instance. If a mechanical system consists of N particles and there are m independent kinematical conditions imposed, it is possible to characterize the system uniquely by n = 3N - m independent generalized coordinates {qk}.[22]

In classical mechanics, the Lagrangian is defined as the kinetic energy,  , of the system minus its potential energy,  .[23] In symbols,

 

Under conditions that are given in Lagrangian mechanics, if the Lagrangian of a system is known, then the equations of motion of the system may be obtained by a direct substitution of the expression for the Lagrangian into the Euler–Lagrange equation, a particular family of partial differential equations.

Here are some definitions:[24]

Definition:
 
is the Lagrange function or lagrangian, qi are the generalized coordinates,   are generalized velocities,
 
are generalized momenta,
 
are generalized forces,
 
are Lagrange's equations.

It is not the purpose here to outline how Lagrangian mechanics works. The interested reader can look at other articles explaining this approach. For the moment, the goal is simply to show that the Lagrangian approach can lead to "generalized fictitious forces" that do not vanish in inertial frames. What is pertinent here is that in the case of a single particle, the Lagrangian approach can be arranged to capture exactly the "coordinate" fictitious forces just introduced.

To proceed, consider a single particle, and introduce the generalized coordinates as {qk} = (r, θ). Then Hildebrand [25] shows in polar coordinates with the qk = (r, θ) the "generalized momenta" are:

 

leading, for example, to the generalized force:

 

with Qr the impressed radial force. The connection between "generalized forces" and Newtonian forces varies with the choice of coordinates. This Lagrangian formulation introduces exactly the "coordinate" form of fictitious forces mentioned above that allows "fictitious" (generalized) forces in inertial frames, for example, the term   Careful reading of Hildebrand shows he doesn't discuss the role of "inertial frames of reference", and in fact, says "[The] presence or absence [of inertia forces] depends, not upon the particular problem at hand but upon the coordinate system chosen." By coordinate system presumably is meant the choice of {qk}. Later he says "If accelerations associated with generalized coordinates are to be of prime interest (as is usually the case), the [nonaccelerational] terms may be conveniently transferred to the right … and considered as additional (generalized) inertia forces. Such inertia forces are often said to be of the Coriolis type."

In short, the emphasis of some authors upon coordinates and their derivatives and their introduction of (generalized) fictitious forces that do not vanish in inertial frames of reference is an outgrowth of the use of generalized coordinates in Lagrangian mechanics. For example, see McQuarrie[26] and Hildebrand.[25] Below is an example of this usage as employed in the design of robotic manipulators:[27] [28][29]

In the above [Lagrange-Euler] equations, there are three types of terms. The first involves the second derivative of the generalized co-ordinates. The second is quadratic in   where the coefficients may depend on  . These are further classified into two types. Terms involving a product of the type   are called centrifugal forces while those involving a product of the type   for i ≠ j are called Coriolis forces. The third type is functions of   only and are called gravitational forces.

— Shuzhi S. Ge, Tong Heng Lee & Christopher John Harris: Adaptive Neural Network Control of Robotic Manipulators, pp. 47-48

The introduction of generalized fictitious forces often is done without notification and without specifying the word "generalized". This sloppy use of terminology leads to endless confusion because these generalized fictitious forces, unlike the standard "state-of-motion" fictitious forces, do not vanish in inertial frames of reference.

Polar coordinates in an inertial frame of reference edit

Below, the "state-of-motion" fictitious forces defined for non-inertial frames are related to the "coordinate" versions for polar coordinates (r, θ). In an inertial frame, let   be the position vector of a moving particle. Its Cartesian components (x, y) are:

 

with r and θ depending on time t.

Unit vectors are defined in the radially outward direction  :

 

and in the direction at right angles to  :

 

These unit vectors vary in direction with time:

 

and:

 

Using these derivatives, the first and second derivatives of position are:

 
 

where dot-overmarkings indicate time differentiation. With this form for the acceleration  , in an inertial frame of reference Newton's second law expressed in polar coordinates is:

 

where F is the net real force on the particle. No fictitious forces appear because all fictitious forces are zero by definition in an inertial frame.

From a mathematical standpoint, however, it sometimes is handy to put only the second-order derivatives on the right side of this equation; that is we write the above equation by rearrangement of terms as:

 

where a "coordinate" version of the "acceleration" is introduced:

 

consisting of only second-order time derivatives of the coordinates r and θ. The terms moved to the force-side of the equation are now treated as extra "fictitious forces" and, confusingly, the resulting forces also are called the "centrifugal" and "Coriolis" force.

These newly defined "forces" are non-zero in an inertial frame, and so certainly are not the same as the previously identified fictitious forces that are zero in an inertial frame and non-zero only in a non-inertial frame.[30] In this article, these newly defined forces are called the "coordinate" centrifugal force and the "coordinate" Coriolis force to separate them from the "state-of-motion" forces.

 
Figure 2: Inertial frame of reference S and instantaneous non-inertial co-rotating frame of reference S' . The co-rotating frame rotates at angular rate Ω equal to the rate of rotation of the particle about the origin of S' at the particular moment t. Particle is located at vector position r(t) and unit vectors are shown in the radial direction to the particle from the origin, and also in the direction of increasing angle θ normal to the radial direction. These unit vectors need not be related to the tangent and normal to the path. Also, the radial distance r need not be related to the radius of curvature of the path.

Co-rotating frame edit

In the case of planar motion of a particle, the "coordinate" centrifugal and Coriolis acceleration terms found above to be non-zero in an inertial frame can be shown to be the negatives of the "state-of-motion" centrifugal and Coriolis terms that appear in a very particular non-inertial co-rotating frame (see next subsection).[31] See Figure 2. To define a co-rotating frame, first an origin is selected from which the distance r(t) to the particle is defined. An axis of rotation is set up that is perpendicular to the plane of motion of the particle, and passing through this origin. Then, at the selected moment t, the rate of rotation of the co-rotating frame Ω is made to match the rate of rotation of the particle about this axis, dθ/dt. The co-rotating frame applies only for a moment, and must be continuously re-selected as the particle moves. For more detail, see Polar coordinates, centrifugal and Coriolis terms.

Polar coordinates in a rotating frame of reference edit

Next, the same approach is used to find the fictitious forces of a (non-inertial) rotating frame. For example, if a rotating polar coordinate system is adopted for use in a rotating frame of observation, both rotating at the same constant counterclockwise rate Ω, we find the equations of motion in this frame as follows: the radial coordinate in the rotating frame is taken as r, but the angle θ' in the rotating frame changes with time:

 

Consequently,

 

Plugging this result into the acceleration using the unit vectors of the previous section:

 
 

The leading two terms are the same form as those in the inertial frame, and they are the only terms if the frame is not rotating, that is, if Ω=0. However, in this rotating frame we have the extra terms:[32]

 

The radial term Ω2 r is the centrifugal force per unit mass due to the system's rotation at rate Ω and the radial term   is the radial component of the Coriolis force per unit mass, where   is the tangential component of the particle velocity as seen in the rotating frame. The term   is the so-called azimuthal component of the Coriolis force per unit mass. In fact, these extra terms can be used to measure Ω and provide a test to see whether or not the frame is rotating, just as explained in the example of rotating identical spheres. If the particle's motion can be described by the observer using Newton's laws of motion without these Ω-dependent terms, the observer is in an inertial frame of reference where Ω=0.

These "extra terms" in the acceleration of the particle are the "state of motion" fictitious forces for this rotating frame, the forces introduced by rotation of the frame at angular rate Ω.[33]

In this rotating frame, what are the "coordinate" fictitious forces? As before, suppose we choose to put only the second-order time derivatives on the right side of Newton's law:

   

If we choose for convenience to treat   as some so-called "acceleration", then the terms   are added to the so-called "fictitious force", which are not "state-of-motion" fictitious forces, but are actually components of force that persist even when Ω=0, that is, they persist even in an inertial frame of reference. Because these extra terms are added, the "coordinate" fictitious force is not the same as the "state-of-motion" fictitious force. Because of these extra terms, the "coordinate" fictitious force is not zero even in an inertial frame of reference.

Notice however, the case of a rotating frame that happens to have the same angular rate as the particle, so that Ω = dθ/dt at some particular moment (that is, the polar coordinates are set up in the instantaneous, non-inertial co-rotating frame of Figure 2). In this case, at this moment, dθ'/dt = 0. In this co-rotating non-inertial frame at this moment the "coordinate" fictitious forces are only those due to the motion of the frame, that is, they are the same as the "state-of-motion" fictitious forces, as discussed in the remarks about the co-rotating frame of Figure 2 in the previous section.

Fictitious forces in curvilinear coordinates edit

 
Figure 3: Coordinate surfaces, coordinate lines, and coordinate axes of general curvilinear coordinates.

Instead of Cartesian coordinates, when equations of motion are expressed in a curvilinear coordinate system, Christoffel symbols appear in the acceleration of a particle expressed in this coordinate system, as described below in more detail. Consider description of a particle motion from the viewpoint of an inertial frame of reference in curvilinear coordinates. Suppose the position of a point P in Cartesian coordinates is (x, y, z) and in curvilinear coordinates is (q1, q2. q3). Then functions exist that relate these descriptions:

  

and so forth. (The number of dimensions may be larger than three.) An important aspect of such coordinate systems is the element of arc length that allows distances to be determined. If the curvilinear coordinates form an orthogonal coordinate system, the element of arc length ds is expressed as:

 

where the quantities hk are called scale factors.[34] A change dqk in qk causes a displacement hk dqk along the coordinate line for qk. At a point P, we place unit vectors ek each tangent to a coordinate line of a variable qk. Then any vector can be expressed in terms of these basis vectors, for example, from an inertial frame of reference, the position vector of a moving particle r located at time t at position P becomes:

 

where qk is the vector dot product of r and ek. The velocity v of a particle at P, can be expressed at P as:

 
 

where vk is the vector dot product of v and ek, and over dots indicate time differentiation. The time derivatives of the basis vectors can be expressed in terms of the scale factors introduced above. for example:

   or, in general,    

in which the coefficients of the unit vectors are the Christoffel symbols for the coordinate system. The general notation and formulas for the Christoffel symbols are:[35][36]

   

and the symbol is zero when all the indices are different. Using relations like this one,

 
 

which allows all the time derivatives to be evaluated. For example, for the velocity:

 
 
 
 

with the Γ-notation for the Christoffel symbols replacing the curly bracket notation. Using the same approach, the acceleration is then

 
 

Looking at the relation for acceleration, the first summation contains the time derivatives of velocity, which would be associated with acceleration if these were Cartesian coordinates, and the second summation (the one with Christoffel symbols) contains terms related to the way the unit vectors change with time.[37]

"State-of-motion" versus "coordinate" fictitious forces edit

Earlier in this article a distinction was introduced between two terminologies, the fictitious forces that vanish in an inertial frame of reference are called in this article the "state-of-motion" fictitious forces and those that originate from differentiation in a particular coordinate system are called "coordinate" fictitious forces. Using the expression for the acceleration above, Newton's law of motion in the inertial frame of reference becomes:

 

where F is the net real force on the particle. No "state-of-motion" fictitious forces are present because the frame is inertial, and "state-of-motion" fictitious forces are zero in an inertial frame, by definition.

The "coordinate" approach to Newton's law above is to retain the second-order time derivatives of the coordinates {qk} as the only terms on the right side of this equation, motivated more by mathematical convenience than by physics. To that end, the force law can be re-written, taking the second summation to the force-side of the equation as:

 

with the convention that the "acceleration"   is now:

 

In the expression above, the summation added to the force-side of the equation now is treated as if added forces were present. These summation terms are customarily called fictitious forces within this "coordinate" approach, although in this inertial frame of reference all "state-of-motion" fictitious forces are identically zero. Thus, the designation of the terms of the summation as "fictitious forces" uses this terminology for contributions that are completely different from the "state-of-motion" fictitious forces. What adds to this confusion is that these "coordinate" fictitious forces are divided into two groups and given the same names as the "state-of-motion" fictitious forces, that is, they are divided into "centrifugal" and "Coriolis" terms, despite their inclusion of terms that are not the "state-of-motion" centrifugal and Coriolis terms. For example, these "coordinate" centrifugal and Coriolis terms can be nonzero even in an inertial frame of reference where the "state-of-motion" centrifugal force (the subject of this article) and Coriolis force always are zero.[38]

If the frame is not inertial, for example, in a rotating frame of reference, the "state-of-motion" fictitious forces are included in the above "coordinate" fictitious force expression.[39] Also, if the "acceleration" expressed in terms of first-order time derivatives of the velocity happens to result in terms that are not simply second-order derivatives of the coordinates {qk} in time, then these terms that are not second-order also are brought to the force-side of the equation and included with the fictitious forces.

Formulation of dynamics in terms of Christoffel symbols and the "coordinate" version of fictitious forces is used often in the design of robots in connection with a Lagrangian formulation of the equations of motion.[29][40]

Notes and references edit

  1. ^ Fictitious forces (also known as a pseudo forces, inertial forces or d'Alembert forces), exist for observers in a non-inertial reference frames. See, for example, "Centrifugal Force". scienceworld.wolfram.com. Retrieved 2008-05-18., "Centrifugal Force - Britannica online encyclopedia". Retrieved 2008-05-18., Max Born & Günther Leibfried (1962). Einstein's Theory of Relativity. New York: Courier Dover Publications. p. pp.76-78. ISBN 0486607690. {{cite book}}: |page= has extra text (help), NASA: Accelerated Frames of Reference: Inertial Forces, Science Joy Wagon: Centrifugal force - the false force
  2. ^ Jerrold E. Marsden, Tudor S. Ratiu (1999). Introduction to Mechanics and Symmetry: A Basic Exposition of Classical Mechanical Systems. Springer. p. p. 251. ISBN 038798643X. {{cite book}}: |page= has extra text (help)
  3. ^ John Robert Taylor (2004). Classical Mechanics. Sausalito CA: University Science Books. p. Chapter 9, pp. 327 ff. ISBN 189138922X.
  4. ^ Stephen T. Thornton & Jerry B. Marion (2004). Classical Dynamics of Particles and Systems (5th Edition ed.). Belmont CA: Brook/Cole. p. Chapter 10. ISBN 0534408966. {{cite book}}: |edition= has extra text (help); line feed character in |author= at position 22 (help)
  5. ^ V. I. Arnol'd (1989). Mathematical Methods of Classical Mechanics. Springer. p. p. 129. ISBN 978-0-387-96890-2. {{cite book}}: |page= has extra text (help)
  6. ^ Harald Iroh (2002). A Modern Approach to Classical Mechanics. World Scientific. p. p. 180. ISBN 9812382135. {{cite book}}: |page= has extra text (help)
  7. ^ K.S. Rao (2003). Classical Mechanics. Orient Longman. p. p. 162. ISBN 8173714363. {{cite book}}: |page= has extra text (help)
  8. ^ Louis N. Hand, Janet D. Finch (1998). Analytical Mechanics. Cambridge University Press. p. p. 267. ISBN 0521575729. {{cite book}}: |page= has extra text (help)
  9. ^ Jean Salençon, Stephen Lyle (2001). Handbook of Continuum Mechanics: General Concepts, Thermoelasticity. Springer. p. p. 9. ISBN 3540414436. {{cite book}}: |page= has extra text (help)
  10. ^ John D. Norton (1993). General covariance and the foundations of general relativity: eight decades of dispute, Rep. Prog. Phys., 56, pp. 835-6.
  11. ^ See Moore and Stommel, Chapter 2, p. 26, which deals with polar coordinates in an inertial frame of reference (what these authors call a "Newtonian frame of reference"), Henry Stommel & Dennis W. Moore (1989). An Introduction to the Coriolis Force. Columbia University Press. p. p. 26. ISBN 0231066368. {{cite book}}: |page= has extra text (help)
  12. ^ For example, Moore and Stommel point our that in a rotating polar coordinate system, the acceleration terms include reference to the rate of rotation of the rotating frame. Henry Stommel & Dennis W. Moore. An Introduction to the Coriolis Force. p. p. 55. {{cite book}}: |page= has extra text (help)
  13. ^ The term particle is used in mechanics to describe an object without reference to its orientation. The term rigid body is used when orientation is also a factor. Thus, the center of mass of a rigid body is a "particle".
  14. ^ Observational frames of reference and coordinate systems are independent ideas. A frame of reference is a physical notion related to the observer's state of motion. A coordinate system is a mathematical description, which can be chosen to suit the observations. A change to a coordinate system that moves in time affects the description of the particle motion, but does not change the observer's state of motion. For more discussion, see Frame of reference
  15. ^ The article on curvature treats a more general case where the curve is parametrized by an arbitrary variable (denoted t), rather than by the arc length s.
  16. ^ Ahmed A. Shabana, Khaled E. Zaazaa, Hiroyuki Sugiyama (2007). Railroad Vehicle Dynamics: A Computational Approach. CRC Press. p. p. 91. ISBN 1420045814. {{cite book}}: |page= has extra text (help)CS1 maint: multiple names: authors list (link)
  17. ^ However, the pilot also will experience Coriolis force, because the pilot is not a particle. When the pilot's head moves, for example, the head has a velocity in the non-inertial frame, and becomes subject to Coriolis force. This force causes pilot disorientation in a turn. See Coriolis effect (perception), Arnauld E. Nicogossian (1996). Space biology and medicine. Reston, VA: American Institute of Aeronautics and Astronautics, Inc. p. p. 337. ISBN 1563471809. {{cite book}}: |page= has extra text (help), and Gilles Clément (2003). Fundamentals of Space Medicine. Springer. p. p. 41. ISBN 1402015984. {{cite book}}: |page= has extra text (help).
  18. ^ Donald T. Greenwood (2003). Advanced Dynamics. Cambridge University Press. p. p. 77. ISBN 0521826128. {{cite book}}: |page= has extra text (help)
  19. ^ Farid M. L. Amirouche (2006). Fundamentals of Multibody Dynamics: Theory and Applications. Springer. p. p. 207. ISBN 0817642366. {{cite book}}: |page= has extra text (help)
  20. ^ Harold Josephs, Ronald L. Huston (2002). Dynamics of Mechanical Systems. CRC Press. p. p. 377. ISBN 0849305934. {{cite book}}: |page= has extra text (help)
  21. ^ Ahmed A. Shabana (2001). Computational Dynamics. Wiley. p. p. 217. ISBN 0471371440. {{cite book}}: |page= has extra text (help)
  22. ^ Cornelius Lanczos (1986). The Variational Principles of Mechanics (Reprint of 4rth Edition of 1970 ed.). Dover Publications. p. p. 10. ISBN 0-486-65067-7. {{cite book}}: |page= has extra text (help)
  23. ^ Cornelius Lanczos (1986). The Variational Principles of Mechanics (Reprint of 4rth Edition of 1970 ed.). Dover Publications. p. pp. 112-113. ISBN 0-486-65067-7. {{cite book}}: |page= has extra text (help)
  24. ^ Vladimir Igorevich Arnolʹd (1989). Mathematical Methods of Classical Mechanics. Springer. p. p. 60. ISBN 0387968903. {{cite book}}: |page= has extra text (help)
  25. ^ a b Francis Begnaud Hildebrand (1965). Methods of Applied Mathematics. Courier Dover Publications. p. pp. 156-157. ISBN 0486670023. {{cite book}}: |page= has extra text (help)
  26. ^ Donald Allan McQuarrie (2000). Statistical Mechanics. University Science Books. p. pp.5-6. ISBN 1891389157. {{cite book}}: |page= has extra text (help)
  27. ^ George F. Corliss, Christele Faure, Andreas Griewank, Laurent Hascoet (editors) (2002). Automatic Differentiation of Algorithms: From Simulation to Optimization. Springer. p. p. 131. ISBN 0387953051. {{cite book}}: |author= has generic name (help); |page= has extra text (help)CS1 maint: multiple names: authors list (link)
  28. ^ Jorge A. C. Ambrósio (editor) (2003). Advances in Computational Multibody Systems. Springer. p. p. 322. ISBN 1402033923. {{cite book}}: |author= has generic name (help); |page= has extra text (help)
  29. ^ a b Shuzhi S. Ge, Tong Heng Lee, Christopher John Harris (1998). Adaptive Neural Network Control of Robotic Manipulators. World Scientific. p. pp. 47-48. ISBN 981023452X. {{cite book}}: |page= has extra text (help)CS1 maint: multiple names: authors list (link)
  30. ^ For a treatment using these terms as fictitious forces, see Henry Stommel, Dennis W. Moore. An Introduction to the Coriolis Force. p. p. 36. ISBN 0231066368. {{cite book}}: |page= has extra text (help)
  31. ^ For the following discussion, see John R Taylor (2005). Classical Mechanics. University Science Books. p. §9.10, pp. 358-359. ISBN 1-891389-22-X. At the chosen instant t0, the frame S' and the particle are rotating at the same rate....In the inertial frame, the forces are simpler (no "fictitious" forces) but the accelerations are more complicated.; in the rotating frame, it is the other way round.
  32. ^ Henry Stommel & Dennis W. Moore (1989). An Introduction to the Coriolis Force. p. p. 55. ISBN 0231066368. {{cite book}}: |page= has extra text (help)
  33. ^ This derivation can be found in Henry Stommel, Dennis W. Moore. An Introduction to the Coriolis Force. p. Chapter III, pp. 54 ff.
  34. ^ PM Morse & H Feshbach (1953). Methods of Mathematical Physics (First Edition ed.). McGraw Hill. p. p. 25. {{cite book}}: |edition= has extra text (help); |page= has extra text (help)
  35. ^ PM Morse & H Feshbach (1953). Methods of Mathematical Physics (First Edition ed.). McGraw Hill. p. pp. 47-48. {{cite book}}: |edition= has extra text (help); |page= has extra text (help)
  36. ^ I-Shih Liu (2002). Continuum mechanics. Springer. p. Appendix A2. ISBN 3540430199.
  37. ^ For application of the Christoffel symbols formalism to a rotating coordinate system, see Ludwik Silberstein (1922). The Theory of General Relativity and Gravitation. D. Van Nostrand. p. pp. 30-32. {{cite book}}: |page= has extra text (help)
  38. ^ For a more extensive criticism of lumping together the two types of fictitious force, see Ludwik Silberstein (1922). The Theory of General Relativity and Gravitation. D. Van Nostrand. p. p. 29. {{cite book}}: |page= has extra text (help)
  39. ^ See Silberstein.
  40. ^ See R. Kelly, V. Santibáñez, Antonio Loría (2005). Control of robot manipulators in joint space. Springer. p. p. 72. ISBN 1852339942. {{cite book}}: |page= has extra text (help)CS1 maint: multiple names: authors list (link)

Further reading edit

External links edit

See also edit