Apparent Forces

Total Page:16

File Type:pdf, Size:1020Kb

Apparent Forces Apparent Forces The motion referred to by Newton’s First Law was for a coordinate system fixed in space (inertial motion). This isn’t valid for motions in our atmosphere since the earth is moving underneath. We must take into account the accelerations of the coordinate reference frame in addition to those of the fluid itself. Apparent Forces are used to include the effects of an accelerating coordinate reference frame (e.g. rotation of the earth). Two apparent forces: Centrifugal force and Coriolis force Centrifugal Force Consider a ball attached to a string or length r, whirling around at a constant angular velocity (ω). From an inertial (fixed coordinate) frame of reference, the speed of the ball is constant but the velocity is changing (since velocity is a vector that includes direction). To calculate the acceleration that the ball experiences, v VD −= ω 2 r Dt This is the expression for the centripetal acceleration. The force causing the acceleration is the string pulling on the ball. Now, if we have the same situation but with the coordinate system rotating with the ball, it would appear to us (stationed on the rotating coordinate system) that the ball would not be moving at all. But there is still a force acting on the ball (pull of the string). So we need to balance the centripetal acceleration by an equal and opposite force. This is called the centrifugal force. The centrifugal force is an apparent force. It is needed to so that we may still use Newton’s laws of motion even with a rotating frame of reference. It’s magnitude is given 2 by Ω RO, where RO is the distance from a point on the earth to the spin axis. Effective Gravity (g) The true gravitational acceleration (g*) pulls all mass towards the center of mass of the earth. However, the centrifugal force (Ω2R) pulls all objects outward from the axis of planetary rotation. The effective gravity (g) is the vector sum of these two forces. It does not point directly at the center of earth mass: Recall that geopotential (Φ) at any point is defined as the work that must be done against gravity in order to raise a mass of 1 kg from sea level to that point. Geopotential by convention = 0 on the surface of the earth. Surfaces of constant geopotential (e.g. the surface of the earth) are normal (perpendicular) to the effective gravity. Since effective gravity does not point directly at the center of the earth’s mass, the constant geopotential surface takes the shape of an oblate ellipsoid instead of a perfect sphere, shown by the red line in the figure below: So the earth “bulges” at the equator. When we use gravity in the equations of motion later on, the centrifugal force remains integrated with g and is not included explicitly. Coriolis Force If an object is moving at the same speed as the earth’s rotation rate (e.g. calm wind), then we only need to invoke the apparent centrifugal force to satisfy Newton’s laws of motion. However, if the object (or air molecules, for instance) are moving faster or slower than the rotation rate of the earth, a second apparent force must be introduced to explain it’s motion. This force is called the Coriolis Force (CF). Note that the earth rotates in a counter-clockwise fashion from a northern hemisphere frame of reference (get a globe, look down from the North Pole, and spin the globe counter clockwise). If we have westerly wind flow (wind from the west), it means that the air is moving faster than the angular velocity of the earth. Therefore, the wind has a higher centrifugal force than that of the earth at that latitude: Ω air = Ωearth + u 2 2 ()earth Ru O Ω>+Ω ROearth If u > 0 and RO is constant So there is a net force acting radially outward from the axis of rotation. In the horizontal plane, there is a southward component (in the northern hemisphere) that acts perpendicular and to the right of the zonal wind, “pulling” the wind to the right. Note that at the Equator, the centrifugal force is perpendicular to the axis of rotation and therefore there is no southward component. Winds do experience a vertical CF at the equator but none in the horizontal plane. The vertical component is very small, however, and need not be considered any further. The same arguments can be made for easterly flow. The wind is moving slower than the angular velocity of the earth, thus the air will have a weaker centrifugal force than that of the earth. This results in a radially inward directed CF, pulling the wind toward the north. Note that this is still perpendicular and to the right of the flow. Through angular momentum arguments and several assumptions (see Holton for complete derivation), the amount of displacement in the zonal, meridional, and vertical directions is given by: Du uv =2 Ωv sinφ + tan φ Dt a Dv u 2 = −2 Ωu sinφ − tan φ Dt a Dw u 2 =2 Ωu cosφ + Dt a Where u = zonal wind, v = meridional wind, w = vertical wind, φ = latitude, and a = radius of earth. The second terms on the right side arise from curvature effects, and can be safely neglected through scale analysis since u << ΩR. Also, the quantity 2Ωsinφ is abbreviated with the letter ‘f’, and is called the Coriolis parameter. It ranges from 0 at the equator to +- 1.46 x 10-4 s-1 at the poles. Thus, for horizontal motion, Du = fv Dt Dv = − fu Dt In vector notation this is written as: CF = -fk x V Where V is the 3-D velocity vector. As we will see later, vertical Coriolis Forces are much smaller than other terms that give rise to vertical motion and will be neglected. In the southern hemisphere, deflection is to the left of the wind. Spin your globe again and turn it upside down to see why. The CF is negligible for motions that have time scales that are much shorter than the period of the earth’s rotation. So you can’t blame it on your bad basketball shooting. The CF does not alter the speed of motion, only the direction. The Coriolis Force is named after French mathematician Gaspard Gustave de Coriolis (1792-1843). .
Recommended publications
  • Rotational Motion (The Dynamics of a Rigid Body)
    University of Nebraska - Lincoln DigitalCommons@University of Nebraska - Lincoln Robert Katz Publications Research Papers in Physics and Astronomy 1-1958 Physics, Chapter 11: Rotational Motion (The Dynamics of a Rigid Body) Henry Semat City College of New York Robert Katz University of Nebraska-Lincoln, [email protected] Follow this and additional works at: https://digitalcommons.unl.edu/physicskatz Part of the Physics Commons Semat, Henry and Katz, Robert, "Physics, Chapter 11: Rotational Motion (The Dynamics of a Rigid Body)" (1958). Robert Katz Publications. 141. https://digitalcommons.unl.edu/physicskatz/141 This Article is brought to you for free and open access by the Research Papers in Physics and Astronomy at DigitalCommons@University of Nebraska - Lincoln. It has been accepted for inclusion in Robert Katz Publications by an authorized administrator of DigitalCommons@University of Nebraska - Lincoln. 11 Rotational Motion (The Dynamics of a Rigid Body) 11-1 Motion about a Fixed Axis The motion of the flywheel of an engine and of a pulley on its axle are examples of an important type of motion of a rigid body, that of the motion of rotation about a fixed axis. Consider the motion of a uniform disk rotat­ ing about a fixed axis passing through its center of gravity C perpendicular to the face of the disk, as shown in Figure 11-1. The motion of this disk may be de­ scribed in terms of the motions of each of its individual particles, but a better way to describe the motion is in terms of the angle through which the disk rotates.
    [Show full text]
  • PHYSICS of ARTIFICIAL GRAVITY Angie Bukley1, William Paloski,2 and Gilles Clément1,3
    Chapter 2 PHYSICS OF ARTIFICIAL GRAVITY Angie Bukley1, William Paloski,2 and Gilles Clément1,3 1 Ohio University, Athens, Ohio, USA 2 NASA Johnson Space Center, Houston, Texas, USA 3 Centre National de la Recherche Scientifique, Toulouse, France This chapter discusses potential technologies for achieving artificial gravity in a space vehicle. We begin with a series of definitions and a general description of the rotational dynamics behind the forces ultimately exerted on the human body during centrifugation, such as gravity level, gravity gradient, and Coriolis force. Human factors considerations and comfort limits associated with a rotating environment are then discussed. Finally, engineering options for designing space vehicles with artificial gravity are presented. Figure 2-01. Artist's concept of one of NASA early (1962) concepts for a manned space station with artificial gravity: a self- inflating 22-m-diameter rotating hexagon. Photo courtesy of NASA. 1 ARTIFICIAL GRAVITY: WHAT IS IT? 1.1 Definition Artificial gravity is defined in this book as the simulation of gravitational forces aboard a space vehicle in free fall (in orbit) or in transit to another planet. Throughout this book, the term artificial gravity is reserved for a spinning spacecraft or a centrifuge within the spacecraft such that a gravity-like force results. One should understand that artificial gravity is not gravity at all. Rather, it is an inertial force that is indistinguishable from normal gravity experience on Earth in terms of its action on any mass. A centrifugal force proportional to the mass that is being accelerated centripetally in a rotating device is experienced rather than a gravitational pull.
    [Show full text]
  • LECTURES 20 - 29 INTRODUCTION to CLASSICAL MECHANICS Prof
    LECTURES 20 - 29 INTRODUCTION TO CLASSICAL MECHANICS Prof. N. Harnew University of Oxford HT 2017 1 OUTLINE : INTRODUCTION TO MECHANICS LECTURES 20-29 20.1 The hyperbolic orbit 20.2 Hyperbolic orbit : the distance of closest approach 20.3 Hyperbolic orbit: the angle of deflection, φ 20.3.1 Method 1 : using impulse 20.3.2 Method 2 : using hyperbola orbit parameters 20.4 Hyperbolic orbit : Rutherford scattering 21.1 NII for system of particles - translation motion 21.1.1 Kinetic energy and the CM 21.2 NII for system of particles - rotational motion 21.2.1 Angular momentum and the CM 21.3 Introduction to Moment of Inertia 21.3.1 Extend the example : J not parallel to ! 21.3.2 Moment of inertia : mass not distributed in a plane 21.3.3 Generalize for rigid bodies 22.1 Moment of inertia tensor 2 22.1.1 Rotation about a principal axis 22.2 Moment of inertia & energy of rotation 22.3 Calculation of moments of inertia 22.3.1 MoI of a thin rectangular plate 22.3.2 MoI of a thin disk perpendicular to plane of disk 22.3.3 MoI of a solid sphere 23.1 Parallel axis theorem 23.1.1 Example : compound pendulum 23.2 Perpendicular axis theorem 23.2.1 Perpendicular axis theorem : example 23.3 Example 1 : solid ball rolling down slope 23.4 Example 2 : where to hit a ball with a cricket bat 23.5 Example 3 : an aircraft landing 24.1 Lagrangian mechanics : Introduction 24.2 Introductory example : the energy method for the E of M 24.3 Becoming familiar with the jargon 24.3.1 Generalised coordinates 24.3.2 Degrees of Freedom 24.3.3 Constraints 3 24.3.4 Configuration
    [Show full text]
  • Chapter 5 ANGULAR MOMENTUM and ROTATIONS
    Chapter 5 ANGULAR MOMENTUM AND ROTATIONS In classical mechanics the total angular momentum L~ of an isolated system about any …xed point is conserved. The existence of a conserved vector L~ associated with such a system is itself a consequence of the fact that the associated Hamiltonian (or Lagrangian) is invariant under rotations, i.e., if the coordinates and momenta of the entire system are rotated “rigidly” about some point, the energy of the system is unchanged and, more importantly, is the same function of the dynamical variables as it was before the rotation. Such a circumstance would not apply, e.g., to a system lying in an externally imposed gravitational …eld pointing in some speci…c direction. Thus, the invariance of an isolated system under rotations ultimately arises from the fact that, in the absence of external …elds of this sort, space is isotropic; it behaves the same way in all directions. Not surprisingly, therefore, in quantum mechanics the individual Cartesian com- ponents Li of the total angular momentum operator L~ of an isolated system are also constants of the motion. The di¤erent components of L~ are not, however, compatible quantum observables. Indeed, as we will see the operators representing the components of angular momentum along di¤erent directions do not generally commute with one an- other. Thus, the vector operator L~ is not, strictly speaking, an observable, since it does not have a complete basis of eigenstates (which would have to be simultaneous eigenstates of all of its non-commuting components). This lack of commutivity often seems, at …rst encounter, as somewhat of a nuisance but, in fact, it intimately re‡ects the underlying structure of the three dimensional space in which we are immersed, and has its source in the fact that rotations in three dimensions about di¤erent axes do not commute with one another.
    [Show full text]
  • The Spin, the Nutation and the Precession of the Earth's Axis Revisited
    The spin, the nutation and the precession of the Earth’s axis revisited The spin, the nutation and the precession of the Earth’s axis revisited: A (numerical) mechanics perspective W. H. M¨uller [email protected] Abstract Mechanical models describing the motion of the Earth’s axis, i.e., its spin, nutation and its precession, have been presented for more than 400 years. Newton himself treated the problem of the precession of the Earth, a.k.a. the precession of the equinoxes, in Liber III, Propositio XXXIX of his Principia [1]. He decomposes the duration of the full precession into a part due to the Sun and another part due to the Moon, to predict a total duration of 26,918 years. This agrees fairly well with the experimentally observed value. However, Newton does not really provide a concise rational derivation of his result. This task was left to Chandrasekhar in Chapter 26 of his annotations to Newton’s book [2] starting from Euler’s equations for the gyroscope and calculating the torques due to the Sun and to the Moon on a tilted spheroidal Earth. These differential equations can be solved approximately in an analytic fashion, yielding Newton’s result. However, they can also be treated numerically by using a Runge-Kutta approach allowing for a study of their general non-linear behavior. This paper will show how and explore the intricacies of the numerical solution. When solving the Euler equations for the aforementioned case numerically it shows that besides the precessional movement of the Earth’s axis there is also a nu- tation present.
    [Show full text]
  • Action Principle for Hydrodynamics and Thermodynamics, Including General, Rotational Flows
    th December 15, 2014 Action Principle for Hydrodynamics and Thermodynamics, including general, rotational flows C. Fronsdal Depart. of Physics and Astronomy, University of California Los Angeles 90095-1547 USA ABSTRACT This paper presents an action principle for hydrodynamics and thermodynamics that includes general, rotational flows, thus responding to a challenge that is more that 100 years old. It has been lifted to the relativistic context and it can be used to provide a suitable source of rotating matter for Einstein’s equation, finally overcoming another challenge of long standing, the problem presented by the Bianchi identity. The new theory is a combination of Eulerian and Lagrangian hydrodynamics, with an extension to thermodynamics. In the first place it is an action principle for adia- batic systems, including the usual conservation laws as well as the Gibbsean variational principle. But it also provides a framework within which dissipation can be introduced in the usual way, by adding a viscosity term to the momentum equation, one of the Euler-Lagrange equations. The rate of the resulting energy dissipation is determined by the equations of motion. It is an ideal framework for the description of quasi-static processes. It is a major development of the Navier-Stokes-Fourier approach, the princi- pal advantage being a hamiltonian structure with a natural concept of energy as a first integral of the motion. Two velocity fields are needed, one the gradient of a scalar potential, the other the time derivative of a vector field (vector potential). Variation of the scalar potential gives the equation of continuity and variation of the vector potential yields the momentum equation.
    [Show full text]
  • Hydraulics Manual Glossary G - 3
    Glossary G - 1 GLOSSARY OF HIGHWAY-RELATED DRAINAGE TERMS (Reprinted from the 1999 edition of the American Association of State Highway and Transportation Officials Model Drainage Manual) G.1 Introduction This Glossary is divided into three parts: · Introduction, · Glossary, and · References. It is not intended that all the terms in this Glossary be rigorously accurate or complete. Realistically, this is impossible. Depending on the circumstance, a particular term may have several meanings; this can never change. The primary purpose of this Glossary is to define the terms found in the Highway Drainage Guidelines and Model Drainage Manual in a manner that makes them easier to interpret and understand. A lesser purpose is to provide a compendium of terms that will be useful for both the novice as well as the more experienced hydraulics engineer. This Glossary may also help those who are unfamiliar with highway drainage design to become more understanding and appreciative of this complex science as well as facilitate communication between the highway hydraulics engineer and others. Where readily available, the source of a definition has been referenced. For clarity or format purposes, cited definitions may have some additional verbiage contained in double brackets [ ]. Conversely, three “dots” (...) are used to indicate where some parts of a cited definition were eliminated. Also, as might be expected, different sources were found to use different hyphenation and terminology practices for the same words. Insignificant changes in this regard were made to some cited references and elsewhere to gain uniformity for the terms contained in this Glossary: as an example, “groundwater” vice “ground-water” or “ground water,” and “cross section area” vice “cross-sectional area.” Cited definitions were taken primarily from two sources: W.B.
    [Show full text]
  • 6. Non-Inertial Frames
    6. Non-Inertial Frames We stated, long ago, that inertial frames provide the setting for Newtonian mechanics. But what if you, one day, find yourself in a frame that is not inertial? For example, suppose that every 24 hours you happen to spin around an axis which is 2500 miles away. What would you feel? Or what if every year you spin around an axis 36 million miles away? Would that have any e↵ect on your everyday life? In this section we will discuss what Newton’s equations of motion look like in non- inertial frames. Just as there are many ways that an animal can be not a dog, so there are many ways in which a reference frame can be non-inertial. Here we will just consider one type: reference frames that rotate. We’ll start with some basic concepts. 6.1 Rotating Frames Let’s start with the inertial frame S drawn in the figure z=z with coordinate axes x, y and z.Ourgoalistounderstand the motion of particles as seen in a non-inertial frame S0, with axes x , y and z , which is rotating with respect to S. 0 0 0 y y We’ll denote the angle between the x-axis of S and the x0- axis of S as ✓.SinceS is rotating, we clearly have ✓ = ✓(t) x 0 0 θ and ✓˙ =0. 6 x Our first task is to find a way to describe the rotation of Figure 31: the axes. For this, we can use the angular velocity vector ! that we introduced in the last section to describe the motion of particles.
    [Show full text]
  • Exploring Artificial Gravity 59
    Exploring Artificial gravity 59 Most science fiction stories require some form of artificial gravity to keep spaceship passengers operating in a normal earth-like environment. As it turns out, weightlessness is a very bad condition for astronauts to work in on long-term flights. It causes bones to lose about 1% of their mass every month. A 30 year old traveler to Mars will come back with the bones of a 60 year old! The only known way to create artificial gravity it to supply a force on an astronaut that produces the same acceleration as on the surface of earth: 9.8 meters/sec2 or 32 feet/sec2. This can be done with bungee chords, body restraints or by spinning the spacecraft fast enough to create enough centrifugal acceleration. Centrifugal acceleration is what you feel when your car ‘takes a curve’ and you are shoved sideways into the car door, or what you feel on a roller coaster as it travels a sharp curve in the tracks. Mathematically we can calculate centrifugal acceleration using the formula: V2 A = ------- R Gravitron ride at an amusement park where V is in meters/sec, R is the radius of the turn in meters, and A is the acceleration in meters/sec2. Let’s see how this works for some common situations! Problem 1 - The Gravitron is a popular amusement park ride. The radius of the wall from the center is 7 meters, and at its maximum speed, it rotates at 24 rotations per minute. What is the speed of rotation in meters/sec, and what is the acceleration that you feel? Problem 2 - On a journey to Mars, one design is to have a section of the spacecraft rotate to simulate gravity.
    [Show full text]
  • Moment of Inertia
    MOMENT OF INERTIA The moment of inertia, also known as the mass moment of inertia, angular mass or rotational inertia, of a rigid body is a quantity that determines the torque needed for a desired angular acceleration about a rotational axis; similar to how mass determines the force needed for a desired acceleration. It depends on the body's mass distribution and the axis chosen, with larger moments requiring more torque to change the body's rotation rate. It is an extensive (additive) property: for a point mass the moment of inertia is simply the mass times the square of the perpendicular distance to the rotation axis. The moment of inertia of a rigid composite system is the sum of the moments of inertia of its component subsystems (all taken about the same axis). Its simplest definition is the second moment of mass with respect to distance from an axis. For bodies constrained to rotate in a plane, only their moment of inertia about an axis perpendicular to the plane, a scalar value, and matters. For bodies free to rotate in three dimensions, their moments can be described by a symmetric 3 × 3 matrix, with a set of mutually perpendicular principal axes for which this matrix is diagonal and torques around the axes act independently of each other. When a body is free to rotate around an axis, torque must be applied to change its angular momentum. The amount of torque needed to cause any given angular acceleration (the rate of change in angular velocity) is proportional to the moment of inertia of the body.
    [Show full text]
  • Newtonian Mechanics Is Most Straightforward in Its Formulation and Is Based on Newton’S Second Law
    CLASSICAL MECHANICS D. A. Garanin September 30, 2015 1 Introduction Mechanics is part of physics studying motion of material bodies or conditions of their equilibrium. The latter is the subject of statics that is important in engineering. General properties of motion of bodies regardless of the source of motion (in particular, the role of constraints) belong to kinematics. Finally, motion caused by forces or interactions is the subject of dynamics, the biggest and most important part of mechanics. Concerning systems studied, mechanics can be divided into mechanics of material points, mechanics of rigid bodies, mechanics of elastic bodies, and mechanics of fluids: hydro- and aerodynamics. At the core of each of these areas of mechanics is the equation of motion, Newton's second law. Mechanics of material points is described by ordinary differential equations (ODE). One can distinguish between mechanics of one or few bodies and mechanics of many-body systems. Mechanics of rigid bodies is also described by ordinary differential equations, including positions and velocities of their centers and the angles defining their orientation. Mechanics of elastic bodies and fluids (that is, mechanics of continuum) is more compli- cated and described by partial differential equation. In many cases mechanics of continuum is coupled to thermodynamics, especially in aerodynamics. The subject of this course are systems described by ODE, including particles and rigid bodies. There are two limitations on classical mechanics. First, speeds of the objects should be much smaller than the speed of light, v c, otherwise it becomes relativistic mechanics. Second, the bodies should have a sufficiently large mass and/or kinetic energy.
    [Show full text]
  • Chapter 04 Rotational Motion
    Chapter 04 Rotational Motion P. J. Grandinetti Chem. 4300 P. J. Grandinetti Chapter 04: Rotational Motion Angular Momentum Angular momentum of particle with respect to origin, O, is given by l⃗ = ⃗r × p⃗ Rate of change of angular momentum is given z by cross product of ⃗r with applied force. p m dl⃗ dp⃗ = ⃗r × = ⃗r × F⃗ = ⃗휏 r dt dt O y Cross product is defined as applied torque, ⃗휏. x Unlike linear momentum, angular momentum depends on origin choice. P. J. Grandinetti Chapter 04: Rotational Motion Conservation of Angular Momentum Consider system of N Particles z m5 m 2 Rate of change of angular momentum is m3 ⃗ ∑N l⃗ ∑N ⃗ m1 dL d 훼 dp훼 = = ⃗r훼 × dt dt dt 훼=1 훼=1 y which becomes m4 x ⃗ ∑N dL ⃗ net = ⃗r훼 × F dt 훼 Total angular momentum is 훼=1 ∑N ∑N ⃗ ⃗ L = l훼 = ⃗r훼 × p⃗훼 훼=1 훼=1 P. J. Grandinetti Chapter 04: Rotational Motion Conservation of Angular Momentum ⃗ ∑N dL ⃗ net = ⃗r훼 × F dt 훼 훼=1 Taking an earlier expression for a system of particles from chapter 1 ∑N ⃗ net ⃗ ext ⃗ F훼 = F훼 + f훼훽 훽=1 훽≠훼 we obtain ⃗ ∑N ∑N ∑N dL ⃗ ext ⃗ = ⃗r훼 × F + ⃗r훼 × f훼훽 dt 훼 훼=1 훼=1 훽=1 훽≠훼 and then obtain 0 > ⃗ ∑N ∑N ∑N dL ⃗ ext ⃗ rd ⃗ ⃗ = ⃗r훼 × F + ⃗r훼 × f훼훽 double sum disappears from Newton’s 3 law (f = *f ) dt 훼 12 21 훼=1 훼=1 훽=1 훽≠훼 P.
    [Show full text]