You are on page 1of 9

1

C4. Kinematics of the material point.1


2. Kinematics of the material point. Simple motions
2.1 Trajectory, velocity, and acceleration in fixed Cartesian system of coordinates
2.2 Rotation vector
2.3 Velocity and acceleration in intrinsic coordinates

2 Kinematics of the material point. Simple motions


Kinematics is a branch of mechanics that describes the motion of mechanical systems
as function of time, without considering cause of the motion, that is without taking into
account mass and forces acting upon systems. In this chapter, we consider the kinematics of
an ideal particle referred to as material point (MP), which has no mass and dimension.
Though choice of coordinate system does not influence the final answer regarding the
characteristics of motion, a wise choice of it may considerably simplify the calculus. The
intrinsic coordinates, the curvilinear orthogonal coordinates are such possible alternative to
the Cartesian coordinate system. In some cases, motion is easier to be described by
introducing relative motion of two coordinate systems. In this chapter, we introduce rotation
vector, and discuss trajectory, velocity and acceleration of the MP for several usual
coordinate systems. Several applications are considered at the end of the next course (C5).

2.1 Trajectory, velocity, and acceleration in fixed Cartesian system of


coordinates
Trajectory is the curve described by MP in motion. Mathematically it is a function
described as a parametric equation, for example, r(t), when time t is the parameter or as an
implicit scalar function of coordinates, for example, f(r)=0.
In Cartesian coordinates, MP moves between the moments t and t '  t  t , and
changes position from r(t) to r(t'). Velocity at moment t is defined as time derivative of the
position vector, that is
r (t ')  r (t ) r (t  t )  r(t ) r(t ) dr(t )
v(t )  lim  lim  lim   r (t ) (2.1.1)
t 't t ' t t 0 t t 0 t dt
or by using Cartesian coordinates
d dx
v  ( xk i k )  k i k  x k i k . (2.1.2)
dt dt
The speed unit in SI is m/s.
Acceleration at moment t is defined as time derivative of velocity, that is
v(t ')  v(t ) v(t  t )  v(t ) v(t ) dv(t )
a(t )  lim  lim  lim 
t ' t t ' t t  0 t t  0 t dt
2
(2.1.3)
d dr (t ) d r (t )
  v (t )   r(t )
dt dt dt 2
or in Cartesian coordinates
d dv
a  ( vk i k )  k i k  vk i k  xk i k . (2.1.4)
dt dt
The acceleration unit in SI is m/s2.
2

2.2 Rotation vector


P1 An infinitesimal rotation can be represented by a vector.
Proof P1
Let us consider a practical problem: make two successive counter-clockwise rotations
of 90o of a six-sided dice about first x and second about z axis, and reverse, first about z and
second about x axis. We obtain a different final orientation of the dice, that is a sequence of
two rotations (a composition of two finite rotation) is not commutative.
Now assume that one takes a general vector c and rotate it about z axis by a small
angle  z (this is called active rotation). This is equivalent with rotating the coordinate axes
about the z axis (this is called passive rotation) by   z .

https://en.wikipedia.org/wiki/Active_and_passive_transformation

Then, with eq. (1.5.11) relation between the rotated and initial coordinate system is
 ex   1   z 0  e x   ex 
      
 ey     z 1 0  e y   [R θ z ] e y  , (2.2.1)
 e   0 
0 1 e z   
 z   ez 
where cos(  z )  1 ,  sin(  z )    z are used. Then, the rotated by  z vector c, that is
the vector c ' is expressed by
 c 'x   1  z 0  cx   cx   cx   z c y 
        
 c ' y     z 1 0  c y   [R θz ]  c y     z cx  c y  . (2.2.2)
c'   0 0 1  c   
 z   cz   z  cz 
Thus, in the first order approximation the rotated vector can be written as
c'  c   z e z  c . (2.2.3)
The above equation can easily be generalized to allow infinitesimal rotations around x and y
axes by  x and  y , respectively. Infinitesimal rotation matrix about x and y axes are
1 0 0  1 0  y 
   
[R θ x ]   0 1   x  , [ R θ y ]   0 1 0. (2.2.4)
 0    
 x 1   y 0 1
One can easily check that the infinitesimal rotation matrices (as they are transformation
matrices they are orthogonal) plus the matrix multiplication operation form an abelian group:
a) closure
[ R θi ][ R θ j ]  [ R θij ] and [ R θij ] is orthogonal ( [ R θij ]1  [ R θij ]T ; prove it);
b) associativity
3

[R θi   
][ R θ j ] [ R θk ]  [ R θi ] [ R θ j ][ R θk ] ;
c) there is an identity orthogonal matrix [ Ι ] (the off-diagonal elements are zero and the
diagonal ones are unity) such as
[ R θi ][ Ι]  [ Ι][ R θi ]  [ R θi ] ;
d) there is an inverse matrix [ R θi ]1 of [ R θi ] such as
[ R θi ][ R θi ]1  [ R θi ]1[ R θi ]  [ Ι] ;
e) the multiplication is commutative
[ R θi ][ R θ j ]  [ R θ j ][ R θi ] ;
i, j, k  x, y,z and the infinitesimal products i  j vanish. Since the vectors and addition
operation form an abelian group, between the vector group and the rotation matrices group
there is an isomorphism defined by the generalization of eq. (2.2.3), that is
c '   R θ c  c  θ  c , (2.2.5)
where the rotation vector is defined by
θ   xe x   y e y   z e z . (2.2.6)
Based on the commutativity of the two groups the above rule of the isomorphism is consistent
since [ R θi ][ R θ j ]  [ R θ j ][ R θi ] and consequently [ R θi ][ R θ j ]c  [ R θ j ][ R θi ]c (according
to eq.(2.2.5)) means
  
c  i ei   j e j  c  c   j e j  i ei  c , 
which holds true only if i are components of a vector (only if i are components of a
vector the commutativity involved by the above equality holds), that is if, generally, θ is a
vector. Consequently, for a general infinitesimal rotation the rotation vector can be introduced
by the relation
c'  c  
θc . (2.2.7)
c
For kinematics, the duration of rotation is relevant. Thus, the time derivative
θ
lim ω , (2.2.8)
t  0 t

defines the angular velocity. From the relation θ  c  c'c  c , with c'  c(t  t ) , we can
write
c(t  t )  c(t ) c θ
c  lim  lim  lim c  ωc , (2.2.9)
t  0 t t  0 t t  0 t

which expresses the time derivative of a vector of constant magnitude.

P2 Vector c of constant magnitude and its time derivative c are perpendicular to each
other.
Proof P2
c  c  c 2  constant .
The time derivative yields
d c  c 
 2c  c  0 ,
dt
that is c  c . A geometrical version of the proof is suggested in Fig. 2.
4

c(t+Δt)

Δc

Δθz
c(t)
Δθ
Δθz θ(t)
z
Δc
c
c c'

O y
ω • c
x

θ
Fig. 1 Infinitesimal rotation vector
about z axis of a constant magnitude Fig. 2 Geometrical construction for
vector. relation c  c .

If we identify by r a rotating position vector of constant magnitude, which is the definition of


circular motion, we have
v  r  ω  r . (2.2.10)
For a general motion ω is generally not perpendicular on r. Its characteristics are discussed in
the following section.

2.3 Velocity and acceleration in intrinsic coordinates


Frenet frame (also called Frenet–Serret) is an intrinsic moving orthogonal frame
whose origin and orientation is function of the current position of MP. To define it, we first
introduce the vector which is tangent to the trajectory. As an application in kinematics, we
end this section by introducing the tangential and normal acceleration.

P3 Velocity is tangent to the trajectory.


Proof P3
To prove P3, we consider the arc length, s(t), as a time dependent coordinate of MP, with the
direction given by the displacement direction of MP along the trajectory. Then
dr (t ) dr ( s(t )) dr ds dr
v(t )     s (2.3.1)
dt dt ds dt ds
We identify t  dr ds as the unit vector tangent to the trajectory (see below) and s the
magnitude of the velocity or the speed (positively defined quantity). Thus, t and v have the
same direction, and algebraically
v (t )  st . (2.3.2)
5

b
t(t)
z
Δt
A: s(t)
B: s(t+Δt)
Δr
P
t’
t’=t(t+Δt)
Δϑ
n
O
r
r+Δr R
Δϑ/2
y

Fig. 3 Frenet frame. t is the tangent unit vector, n is


the normal unit vector, and b is the binormal unit
vector. The osculating circle of radius of curvature R
is the plane of t and n.

A geometrical description might help for an easier understanding of the notions introduced
above. Thus, in Fig. 3, we represented the trajectory of point P in the three dimensional
Euclidean space (trajectory is not restricted to an in-plane curve). An infinitesimal
displacement along arc AB, between s(t) and s(t+Δt), which is a portion of the osculating
circle, is drawn. The osculating circle is the circle of maximum contact with the trajectory at
any point. The osculating circle coincides with the circular trajectory in a circular motion. Its
radius is the radius of curvature. The inverse of the radius of curvature is called curvature.
With the geometry of Fig. 3, we can write
r r s dr ds
v  lim  lim  , (2.3.3)
t 0 t t 0 s t ds dt
and

r 2 R sin
lim  lim 2  1, (2.3.4)
t 0 s  0 
2R
2
where R is the radius of curvature. To make connection with the tangential and normal
acceleration, we consider  a positively defined quantity. By inspection, we observe
r dr
lim  t, (2.3.5)
s0 s ds
where t is the tangent vector of magnitude unity. On another hand
s
lim  s , (2.3.6)
t 0 t
6

which is positively defined (the arc length is a positive quantity which increases in time) and
we obtain
v  st . (2.3.7)
-------------------------------------------------------------------
Next, we introduce the Frenet frame. The motion of point P along a three-dimensional
path is considered in Fig. 3. The tangent direction is defined by the unit tangent vector t. By
differentiating t  t  1 with respect to s coordinate, we obtain
dt
t   0,
ds
that is dt ds is a vector perpendicular to the vector t. Geometrically, for the tangent vector
derivative with respect to  , we can write (see Fig. 3)

dt 2 t sin
 lim 2  1, ` (2.3.8)
d  0 2 
2
Regarding the direction, we observe that t  tends to be an inward curve vector (oriented
to the interior of the curve) perpendicular to t (this holds since  is a positively defined
quantity). The unit vector with this orientation is referred to as the unit normal vector, and is
defined by the relation
dt t
 lim n . (2.3.9)
d   0 
Then we can write
dt dt d n
  , (2.3.10a)
ds d ds R
dt dt ds n
  s , (2.3.10b)
dt ds dt R
where the definition for the radius of curvature is introduced by
d 1
  , (2.3.11)
ds R
and  is curvature. Eq. (2.3.10a) is the first Frenet formula. As ds and d are positive R is
positively defined. Such positive R characterizes the so-called regular curves. A more general
treatment, which uses both positive and negative (or zero) curvatures is a subject of
differential geometry. As connection with kinematics is much more involving when the
general geometrical description of the curve is important, we only give some references on
this subject (see Further reading section).
In addition to the vectors t and n, the unit binormal vector b is introduced by relation
(see Fig. 3)
b  tn . (2.3.12)
The Frenet frame is moving along the curve and has a rotating motion. The instantaneous
rotation vector ω D associated to the time variation of the unit vectors t, b, n is (often) called
Darboux vector. It can be obtained in terms of the Frenet characteristics of motion by using
eq. (2.2.9) as follows. The rotation vector is generally written as
ω D  t t  n n  bb (2.3.13)
and with eq. (2.3.10b)
t  ωD  t  n b  b n  s n (2.3.14)
R
that is n  0 and b  s R  s . Also
7

 t b
n  ωD  n  t b  b t  s      s   t   b  , (2.3.15)
 R T
where T is the radius of torsion, and  is torsion, and t  s T  s .
n
b  ωD  b  n t  t n   s   sn (2.3.16)
T
By inspection of eqs. (2.1.13-16) one obtains
s s
ωD  t  b . (2.3.17)
T R
For a plane curve, T   and
s
ωD  b (2.3.18)
R
and as both s and R are positive quantities ω has direction of b. Intuitively, the curvature
measures the failure of a curve to be a straight line, while torsion measures the failure of a
curve to be in-plane.

Exercise (2.3.1)
Prove the following relations (Frenet formulae):
dt n
i)  s  n ;
ds R
dn t b
ii)     t  b ;
ds R T
db n
iii)    n .
ds T

Exercise (2.3.2)
Prove the following relations:
r t r  r
i) t  ; ii) n  ; iii) b  ;
r t r  r
r  r r  r  r
iv)  ; v)   .
r
3
r  r  r  r
---------------------------------------------------------------------------------------------------------

Tangential and normal acceleration


s 2
P4 The acceleration vector is sum of tangential, a t  st , and normal, a n  n,
R
accelerations
a  at  an . (2.3.19)
Proof P4
To prove P4, we take the time derivative of eq. (2.3.7),
a  st  st (2.3.20)
Time derivative of the second term, with the chain rule, can be written as with eq. (2.3.10b) as
dt ds 1
t   s n , (2.3.21)
ds dt R
--------------------------------------------------------
8

Comment
With Fig. 4 one can obtain a more intuitive picture about the definition of curvature. The
angle  in Fig. 4 is the angle of slope of the curve (if y  y ( x ) , then tan   dy / dx ). Thus,
we notice that  (t )   / 2   (t ) ,  (t  t )   / 2   (t  t ) ,  (t  t )   (t )   , and one
obtains that    . Generally, d    d , and
dt dt d  dt d 1
   n (2.3.22)
ds d  ds d ds R
With the adopted convention of the positive sign for the curvature we may calculate the radius
of curvature by the derivative of slope angle as
d d 1
  (2.3.23)
ds ds R

Δs
Δϑ

R
α(t+ Δt) γ (t)
α(t)
γ'= γ(t+Δt) x

Fig. 4 For in-plane curve,


s  R and Δs,  , R are
positively defined.    '  is
equal to   as function of
velocity direction on the curve (here
   ).

--------------------------------------------------------

Concluding, for eq. (2.3.19), we obtain proposition P4


s 2
a  st  n , (2.3.24)
R
where at  
st , and an  s 2 R 1n .

Exercise (2.3.3)
Show that for an in-plane curve y  y ( x ) , the radius of curvature R is given by

1 y 
 
1  y  
3
R 2 2

In Fig. 5, the tangential and normal accelerations are represented for a three
dimensional Euclidean space trajectory. Tangential acceleration is parallel to the velocity, in
9

the same direction for increasing speed, and in opposite direction for decreasing speed. It
characterizes speed change. Normal acceleration has direction of the normal vector, that is,
inward oriented (to the interior of the trajectory) and perpendicular to the trajectory tangent
vector. It characterizes velocity orientation change and is not vanishing even in the case the
speed is constant but velocity orientation changes.

z b

n t
at
an

y
a
x

Fig. 5 Normal a n and tangential a t


accelerations.

P5 The normal acceleration can be written as an  ω D  v , where ω D is the Darboux


vector associated to the instantaneous rotation of the intrinsic coordinate system.
Proof P5
From eqs. (2.3.19, 20), by replacing c by t in eq. (2.2.9), we have,
an  st  sω D  t  ω D  v (2.3.25)

You might also like