Download as pdf or txt
Download as pdf or txt
You are on page 1of 53

Predicting Outdoor Sound 2nd Edition

Kai Ming Li
Visit to download the full and correct content document:
https://textbookfull.com/product/predicting-outdoor-sound-2nd-edition-kai-ming-li/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Proceedings of the 6th Conference on Sound and Music


Technology CSMT Revised Selected Papers Wei Li

https://textbookfull.com/product/proceedings-of-the-6th-
conference-on-sound-and-music-technology-csmt-revised-selected-
papers-wei-li/

Proceedings of the 7th Conference on Sound and Music


Technology CSMT Revised Selected Papers Haifeng Li

https://textbookfull.com/product/proceedings-of-the-7th-
conference-on-sound-and-music-technology-csmt-revised-selected-
papers-haifeng-li/

Acoustics Sound Fields Transducers and Vibration 2nd


Edition Leo Beranek

https://textbookfull.com/product/acoustics-sound-fields-
transducers-and-vibration-2nd-edition-leo-beranek/

Privacy Computing Theory and Technology 2nd Edition


Fenghua Li

https://textbookfull.com/product/privacy-computing-theory-and-
technology-2nd-edition-fenghua-li/
Animal Rights Education Kai Horsthemke

https://textbookfull.com/product/animal-rights-education-kai-
horsthemke/

Nols Cookery National Outdoor Leadership School

https://textbookfull.com/product/nols-cookery-national-outdoor-
leadership-school/

Musical Sound Effects Analog and Digital Sound


Processing 1st Edition Reveillac

https://textbookfull.com/product/musical-sound-effects-analog-
and-digital-sound-processing-1st-edition-reveillac/

Predicting Movie Success at the Box Office 1st Edition


Barrie Gunter

https://textbookfull.com/product/predicting-movie-success-at-the-
box-office-1st-edition-barrie-gunter/

Advances in psychopharmacology : predicting and


improving treatment response First Edition Carman

https://textbookfull.com/product/advances-in-psychopharmacology-
predicting-and-improving-treatment-response-first-edition-carman/
Predicting Outdoor Sound
r.?\ Taylor & Francis
� Taylor & Francis Group
http://taylorandfrancis.com
Predicting Outdoor Sound
Second Edition

Keith Attenborough
and
Timothy Van Renterghem
Second edition published 2021
by CRC Press
6000 Broken Sound Parkway NW, Suite 300, Boca Raton, FL 33487-2742

and by CRC Press


2 Park Square, Milton Park, Abingdon, Oxon, OX14 4RN

© 2021 Keith Attenborough and Timothy Van Renterghem

First edition published by CRC Press 2007

CRC Press is an imprint of Taylor & Francis Group, LLC

The right of Keith Attenborough and Timothy Van Renterghem to be identified as


authors of this work has been asserted by them in accordance with sections 77 and 78
of the Copyright, Designs and Patents Act 1988.

Reasonable efforts have been made to publish reliable data and information, but the
author and publisher cannot assume responsibility for the validity of all materials or
the consequences of their use. The authors and publishers have attempted to trace the
copyright holders of all material reproduced in this publication and apologize to copy-
right holders if permission to publish in this form has not been obtained. If any copy-
right material has not been acknowledged please write and let us know so we may
rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted,
reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other
means, now known or hereafter invented, including photocopying, microfilming, and
recording, or in any information storage or retrieval system, without written permission
from the publishers.

For permission to photocopy or use material electronically from this work, access
www.copyright.com or contact the Copyright Clearance Center, Inc. (CCC), 222
Rosewood Drive, Danvers, MA 01923, 978-750-8400. For works that are not avail-
able on CCC please contact mpkbookspermissions@tandf.co.uk

Trademark notice: Product or corporate names may be trademarks or registered trade-


marks and are used only for identification and explanation without intent to infringe.

ISBN: 978-1-4987-4007-4 (hbk)


ISBN: 978-0-4294-7080-6 (ebk)

Typeset in Sabon
by SPi Global, India
Contents

Preface xv
Authors Biography xvii

1 Introduction 1
1.1 Early observations 1
1.2 A brief survey of outdoor sound attenuation mechanisms 2
1.3 Data illustrating ground effect 3
1.3.1 Propagation from a fixed jet engine source 3
1.3.2 Propagation over discontinuous ground 5
1.4 Data illustrating the combined effects of
ground and meteorology 8
1.4.1 More fixed jet engine data 8
1.4.2 Road traffic noise propagation over flat terrain
under strong temperature inversion 9
1.4.3 Meteorological effects on railway noise
propagation over flat terrain 14
1.4.4 Road traffic noise propagation in a valley 18
1.5 Classification of meteorological conditions
for outdoor sound prediction 21
1.6 Typical sound speed profiles 29
1.7 Linear-logarithmic representations of sound speed profiles 34
1.8 Air absorption 40
Note 43
References 43

2 The propagation of sound near ground surfaces in a


homogeneous medium 45
2.1 Introduction 45
2.2 A point source above smooth flat acoustically soft ground 45
2.3 The sound field above a locally reacting ground 51
2.4 The sound field above a layered extended-reaction ground 57

v
vi Contents

2.5 Surface waves above porous ground 64


2.6 Experimental data and numerical predictions 67
2.7 The sound field due to a line source near the ground 72
References 75

3 Predicting effects of source characteristics 79


3.1 Introduction 79
3.2 Sound fields due to dipole sources near the ground 79
3.2.1 The horizontal dipole 80
3.2.2 The vertical dipole 86
3.2.3 An arbitrarily orientated dipole 91
3.3 The sound field due to an arbitrarily orientated quadrupole 96
3.4 Railway noise directivity and prediction 101
3.5 Source characteristics of road traffic 103
3.5.1 Basic formulae and parameters 103
3.5.2 Directivity corrections 107
3.5.3 Other corrections and limitations 109
3.6 Source characteristics of wind turbines 111
3.6.1 Sound-generation mechanisms 111
3.6.2 Typical spectra of large horizontal axis wind turbines 112
3.6.3 Horizontal and vertical directivity 113
3.6.4 Amplitude modulation 114
References 116

4 Numerical methods based on time-domain approaches 119


4.1 Introduction 119
4.2 An efficient complete finite-difference time-domain
model for outdoor sound propagation 120
4.2.1 Sound propagation equations 120
4.2.2 Numerical discretization 122
4.2.2.1 Homogeneous and still propagation medium 123
4.2.2.2 Inhomogeneous media 125
4.2.2.3 Moving medium 126
4.2.2.4 Numerical accuracy and stability 128
4.2.3 Modelling propagation in a moving unbounded
atmosphere 130
4.2.4 Modelling finite impedance boundary conditions 132
4.2.4.1 Impedance plane approach 132
4.2.4.2 Ground interaction modelling by including
a layer of soil 134
4.3 Long distance sound propagation prediction based on FDTD 136
4.3.1 Moving frame FDTD 136
Contents vii

4.3.2 Hybrid modelling: combining FDTD with GFPE 137


4.3.2.1 Advantages of the GFPE method 138
4.3.2.2 Complex source region, simplified receiver
region 139
4.3.2.3 Procedure for one-way coupling from FDTD
to GFPE 139
4.3.2.4 Numerical example 140
4.3.2.5 Computational cost reduction 142
References 143

5 Predicting the acoustical properties of ground surfaces 147


5.1 Introduction 147
5.2 Predicting ground impedance 148
5.2.1 Empirical and phenomenological models 148
5.2.2 Microstructural models using idealized pore shapes 151
5.2.3 Approximate models for high flow resistivities 160
5.2.4 Relaxation models 163
5.2.5 Relative influence of microstructural parameters 165
5.3 Physical inadmissibility of semi-empirical models 168
5.4 Predicting effects of surface roughness 171
5.4.1 Boss and stochastic models 171
5.4.2 Impedance models including rough surface effects 175
5.4.2.1 Hard rough surfaces 175
5.4.2.2 Rough finite impedance surfaces 180
5.4.2.3 Modified ‘boss’ and empirical models for
regularly spaced roughness elements 187
5.4.2.4 Multiple scattering models 187
5.4.2.5 A roughness spectrum model 194
5.4.3 Propagation over rough seas 194
5.4.3.1 Effective impedance of rough sea surfaces 194
5.4.3.2 Predicted propagation of
offshore pilling noise 198
5.4.3.3 Predicted rough sea effects on sonic booms 201
5.5 Predicting effects of ground elasticity 202
5.5.1 Coupling from airborne sound to structures and ground
vibration 202
5.5.2 Biot-Stoll theory 203
5.5.3 Numerical calculations of acoustic-seismic coupling 206
5.5.3.1 Fast field program for layered air-ground
systems (FFLAGS) 206
5.5.3.2 Example predictions of low-frequency effects 209
References 215
viii Contents

6 Measurements of the acoustical properties of ground


surfaces and comparisons with models 221
6.1 Impedance measurement methods 221
6.1.1 Impedance tube 221
6.1.2 Impedance meter 222
6.1.3 Non-invasive measurements 222
6.1.3.1 Direct measurement of reflection coefficient 222
6.1.3.2 Impedance deduction from short-range
measurements 224
6.1.3.3 Model parameter deduction from short-range
propagation data 227
6.1.3.4 A template method for
impedance deduction 228
6.1.3.5 Effective flow resistivity classification 230
6.1.3.6 Direct impedance deduction 230
6.2 Comparisons of impedance spectra with model
predictions 232
6.3 Fits to short-range propagation data using impedance
models 234
6.3.1 Short-range grassland data and fits 234
6.3.2 Fits to data obtained over forest floors, gravel and
porous asphalt 238
6.3.3 Railway ballast 243
6.3.4 Measured flow resistivities and porosities 246
6.3.5 Comparison of template and direct deduction
methods over grassland 247
6.4 Spatial and seasonal variations in grassland impedance 248
6.4.1 Predicted effects of spatial variation 248
6.4.2 Measured effects of varying moisture content 250
6.4.3 Influence of water content on ‘fast’ and shear
wave speeds 251
6.4.4 Measured spatial and seasonal variations 252
6.5 Ground effect predictions based on fits to short-range
level difference spectra 253
6.6 On the choice of ground impedance models
for outdoor sound prediction 258
6.7 Measured and predicted surface roughness effects 261
6.7.1 Roughness-induced ground effect 261
6.7.2 Excess attenuation spectra for random and periodic
roughness 262
6.7.3 Roughness-induced surface waves 267
6.7.4 Outdoor measurements of the influence of roughness
on ground effect 277
Contents ix

6.8 Measured and predicted effects of ground elasticity 280


6.8.1 Elasticity effects on surface impedance 280
6.8.2 Ground vibrations due to airborne explosions 281
6.9 Non-linear interaction with porous ground 292
6.10 Deduction of soil properties from measurement of
A/S coupling 293
References 298

7 Influence of source motion on ground effect and diffraction 305


7.1 Introduction 305
7.2 A monopole source moving at constant speed and height above
a ground surface 306
7.3 The sound field of a source moving with arbitrary velocity 312
7.4 Comparison with heuristic calculations 318
7.5 Point source moving at constant speed and height parallel to
a rigid wedge 319
7.5.1 Kinematics 319
7.5.2 Diffracted pressure for a source in uniform
motion 322
7.6 Source moving parallel to a impedance discontinuity 325
7.6.1 Introduction 325
7.6.2 Uniform motion parallel to a single discontinuity 327
7.7 Source moving at constant height parallel to a rigid barrier
above the ground 331
7.7.1 Barrier over hard ground 331
7.7.2 Barrier over impedance ground 334
7.8 Source moving over externally reacting ground 336
References 339

8 Predicting effects of mixed impedance ground 341


8.1 Introduction 341
8.2 Single impedance discontinuity 342
8.2.1 De Jong’s semi-empirical method 342
8.2.2 Modified De Jong method 343
8.2.3 Rasmussen’s method 344
8.3 Multiple impedance discontinuities 345
8.3.1 An extended De Jong method 345
8.3.2 The nMID (multiple impedance discontinuities) method 346
8.3.3 Nyberg’s method 347
8.3.4 Fresnel-zone methods 348
8.3.5 The boundary element method 352
8.4 Comparisons of predictions with data 356
8.4.1 Single impedance discontinuity 356
8.4.2 Impedance strips 357
x Contents

8.5 Refraction above mixed impedance ground 361


8.6 Predicting effects of ground treatments near surface transport 365
8.6.1 Roads 365
8.6.1.1 Sound propagation from a road over
discontinuous impedance 365
8.6.1.2 Predicted effects of replacing ‘Hard’
by ‘Soft’ ground near a road 367
8.6.1.3 Predicting effects of low parallel
walls and lattices 370
8.6.2 Tramways 371
8.6.3 Railways 372
8.6.3.1 Porous sleepers and porous slab track 372
8.7 Predicting meteorological effects on the insertion loss of
low parallel walls 379
8.7.1 Configuration and geometry 379
8.7.2 Numerical methods 379
8.7.3 Meteorological effects 380
8.8 Predicting effects of variability in downward-refraction and
ground impedance 382
8.8.1 Introduction 382
8.8.2 Meteorological data and processing 383
8.8.3 Grassland impedance data 385
8.8.4 Sound propagation modelling and
numerical parameters 385
8.8.5 Detailed analysis of a temporal sequence 386
8.8.6 Statistical analysis of temporal variation over a
full year 389
8.8.6.1 Spectral variation 389
8.8.6.2 Variation in A-weighted pink noise 390
8.8.6.3 Convergence to yearly LAeq 392
8.8.6.4 Conclusions 393
References 394

9 Predicting the performance of outdoor noise barriers 397


9.1 Introduction 397
9.2 Analytical solutions for the diffraction of sound by a barrier 398
9.2.1 Formulation of the problem 398
9.2.2 The MacDonald solution 401
9.2.3 The Hadden and Pierce solution for a wedge 404
9.2.4 Approximate analytical formulation 407
9.3 Empirical formulations for studying the shielding effect
of barriers 411
9.4 The sound attenuation by a thin plane on the ground 416
9.5 Noise reduction by a finite-length barrier 420
9.6 Adverse effect of gaps in barriers 423
Contents xi

9.7 The acoustic performance of an absorptive screen 429


9.8 Gabion barriers 432
9.8.1 Numerical predictions of comparative
acoustical performance 432
9.8.2 Laboratory measurements on porous-stone gabions 435
9.8.3 Outdoor measurements on a gabion barrier 438
9.8.4 Optimizing gabion barriers for noise reduction 438
9.9 Other factors in barrier performance 439
9.9.1 Barrier shape 439
9.9.2 Meteorological effects on barrier performance 444
9.9.3 Rough and soft berms 446
9.9.4 Berms vs barriers in wind 448
9.10 Sonic crystal noise barriers 452
9.11 Predicted effects of spectral variations in train noise during pass-by 455
References 459

10 Predicting effects of vegetation, trees and turbulence 467


10.1 Measured effects of vegetation 467
10.1.1 Influence of vegetation on soil properties 467
10.1.2 Measurements of sound transmission through vegetation 471
10.1.3 Measured attenuation due to trees,
shrubs and hedges 474
10.2 Predicting sound transmission through vegetation 477
10.2.1 Ground effect with plants and vegetation 477
10.2.2 Models for foliage effects 480
10.2.2.1 Empirical models 480
10.2.2.2 Scattering models 484
10.2.3 Reduction of coherence by scattering 487
10.2.4 Predictions of ground effect, scattering
and foliage attenuation 490
10.2.4.1 Sound propagation in crops 490
10.2.4.2 Sound propagation in forests 495
10.3 Influence of ground on propagation through
arrays of vertical cylinders 499
10.3.1 Laboratory data combining ‘Sonic
Crystal’ and ground effects 499
10.3.2 Numerical design of tree belts for traffic noise reduction 502
10.3.3 Measured and predicted effects of irregular
spacing in the laboratory 506
10.4 Reflection from forest edges 508
10.5 Meteorological effects on sound transmission through trees 511
10.6 Combined effects of trees, barriers and meteorology 516
10.7 Turbulence and its effects 520
10.7.1 Turbulence mechanisms 520
10.7.2 Models for turbulence spectra 522
xii Contents

10.7.3 Clifford and lataitis prediction of ground


effect in turbulent conditions 525
10.7.4 Ostashev et al. improvements on the
Clifford and lataitis approach 526
10.7.5 Height dependence of turbulence 529
10.7.6 Turbulence-induced phase and log-
amplitude fluctuations 530
10.7.7 Scattering by turbulence 531
10.7.8 Decrease in sound levels due to turbulence 531
10.7.9 Measurement of turbulence 532
10.7.10 Inclusion of atmospheric turbulence
in the fast field program 533
10.7.11 Comparisons with experimental data 534
10.7.12 Including turbulence in FDTD calculations 536
10.8 Equivalence of turbulence and scattering
influences on coherence 539
References 542

11 Ray tracing, analytical and semi-empirical


approximations for a-weighted levels 549
11.1 Ray tracing 549
11.2 Linear sound speed gradients and weak refraction 558
11.3 Approximations for A-weighted levels and
ground effect optimization in the presence of
weak refraction and turbulence 561
11.3.1 Ground effect optimization 561
11.3.2 Integral expressions for A-weighted
mean square sound pressure 561
11.3.3 Approximate models for ground impedance 564
11.3.4 Effects of weak refraction 564
11.3.5 Approximations for excess attenuation 565
11.3.5.1 Variable porosity or thin layer ground 565
11.3.5.2 Rough ground 566
11.3.5.3 Smooth high flow resistivity ground 568
11.3.6 Numerical examples and discussion 568
11.3.6.1 Comparison with data: Avon
jet engine source 568
11.3.6.2 Sensitivity to spectrum, source
height and distance 568
11.3.6.3 Variation with distance 573
11.3.6.4 Effects of refraction 575
11.3.7 Concluding remarks 576
11.4 A semi-empirical model for A-weighted sound
levels at long range 578
References 580
Contents xiii

12 Engineering models 583


12.1 Introduction 583
12.2 ISO 9613–2 583
12.2.1 Description 583
12.2.2 Basic equations 584
12.2.2.1 Geometrical divergence 586
12.2.2.2 Atmospheric absorption 586
12.2.2.3 Ground effect 586
12.2.2.4 Screening 587
12.2.2.5 Meteorological correction 588
12.2.3 General critique 589
12.2.4 Accuracy of ISO 9613-2 ground effect 590
12.3 CONCAWE 593
12.3.1 Introduction 593
12.3.2 Basis and provisions of scheme 593
12.3.3 Criticisms of CONCAWE 595
12.4 Calculation of road traffic noise (CRTN) 596
12.4.1 Introduction 596
12.4.2 Basic equations 598
12.4.2.1 L10 levels 598
12.4.2.2 Corrections for mean traffic speed, percentage
of heavy vehicles and gradient 600
12.4.2.3 Correction for type of road surface 601
12.4.2.4 Distance correction 601
12.4.2.5 Ground cover correction 601
12.4.2.6 Screening correction 602
12.4.2.7 Site layout 604
12.4.2.8 Segments and road junctions 605
12.5 Calculation of railway noise (CRN) 605
12.6 NORD2000 607
12.7 HARMONOISE 607
12.7.1 Introduction and background 607
12.7.2 General methodology 608
12.7.2.1 Basic equations 608
12.7.2.2 Identification of propagation planes 610
12.7.2.3 Recommended numerical techniques 610
12.7.2.4 Meteorological conditions 611
12.7.2.5 Frequency resolution 611
12.7.2.6 Long-term integrated levels 611
12.7.2.7 Validation 612
12.7.3 Analytical point-to-point model 613
12.7.3.1 Introduction 613
12.7.3.2 Methodology for combining
ground and barrier effect 613
12.7.3.3 Ground reflection model 614
xiv Contents

12.7.3.4 Sound diffraction model 616


12.7.3.5 Transition model 618
12.7.3.6 Refraction 618
12.7.3.7 Coherence losses 621
12.7.3.8 Scattering by turbulence 622
12.8 The Environmental noise directive (END) scheme
(CNOSSOS-EU) 622
12.8.1 Ground effect 622
12.8.2 Criticisms 625
12.9 Performance of railway noise prediction schemes in
high-rise cities 626
12.10 Performance of engineering models in a complex road traffic
noise example 633
12.10.1 Site and models 633
12.10.2 Approximating the berm slope 634
12.10.3 Road traffic source power modelling 635
12.10.4 Daytime vs nighttime measurements and
predictions 636
12.10.5 Model performance 636
12.11 Predicting wind turbine noise 639
12.11.1 An untypical industrial source 639
12.11.2 Complex meteorologically induced
propagation effects 639
12.11.3 Ground effect for wind turbine sound propagation 640
12.11.4 Propagation over non-flat terrain 642
12.12 Prediction requirements for outdoor sound auralization 643
12.12.1 Introduction 643
12.12.2 Simulating outdoor attenuation by filters 644
12.12.3 Auralization of a noise abatement based
on a priori recordings 645
References 646

Index����������������������������������������������������������������������������������������������� 653
Preface

Outdoor sound propagation is of wide-ranging interest not only for predicting


noise exposure but also in animal bioacoustics and in military contexts. Based
on wide-ranging backgrounds in research and consultancy, the 2nd edition of
‘Predicting Outdoor Sound’ aims to provide a comprehensive reference on
aspects of outdoor sound propagation and its prediction that should be useful
to practitioners yet is respectable from the academic point of view. Despite the
significant progress in theories and, particularly, in numerical methods for the
various phenomena that are involved in outdoor sound propagation since the
1st edition of this text was published in 2006, current prediction schemes for
outdoor sound remain largely empirical. While empirical approaches are under-
standable in view of the complicated source characteristics and complex propa-
gation paths that are often of interest, numerical methods and theories have
been validated extensively by comparisons with data and help with our under-
standing of the important effects. This text aims to bring the leading theories
and data together and to provide the noise consultant or relevant practitioner
with the basis for deciding between models and schemes for use in any given
situation. Enough detail is presented to make the reader aware of the inherent
approximations, restrictions and difficulties of the prediction methods
discussed.
The text does not attempt to duplicate the comprehensive treatments of
theoretical and numerical models in “Computational Atmospheric
Acoustics” by Erik Salomons published in 2001 and in “Acoustics in Moving
Inhomogeneous Media” by Vladimir Ostashev and Keith Wilson (2nd edi-
tion, a paperback version was published in 2019). While these texts are
excellent, neither of them includes any data. In contrast, this text emphasizes
data and introduces aspects of research not mentioned in these texts. Those
interested in outdoor sound prediction would benefit from all three texts.
I am delighted that Timothy Van Renterghem has helped to produce this
2nd edition. He is well-known internationally for his work on many aspects
of outdoor sound propagation. We worked together on an EC FP7 project1
“Holistic and sustainable abatement of noise by optimized combinations of
natural and artificial means” (HOSANNA) and it has been a pleasure to
continue working with him. Some of the material in this second edition

xv
xvi Preface

results from the HOSANNA project. The additions include comparisons


between predictions and data for noise from road traffic, railways and wind
turbines (Chapters 1 and 12), extended descriptions of the modelling of
source characteristics, including the HARMONOISE model and its propa-
gation modules (Chapters 3 and 12), predictions of propagation over rough
seas, parallel low walls and lattices (Chapters 5, 8 and 9), descriptions of
numerical methods (chapter 4), gabion and sonic crystal noise barriers per-
formance and design (chapter 9), meteorological effects on noise barrier
performance (Chapter 9), numerical designs of tree belts for road traffic
noise reduction (Chapter 10) and requirements for auralization (Chapter
12). I would like to acknowledge the support of my previous co-authors, Kai
Ming Li and Kirill Horoshenkov. Most of their contributions to the 1st edi-
tion are retained, except that, given the imminent appearance of a book by
Maarten Hornikx and Timothy Van Renterghem on Urban Sound
Propagation, the chapter devoted to this topic has been removed.

Keith Attenborough
June 2020

NOTE

1 https://cordis.europa.eu/project/id/234306
Authors’ Biographies

Keith Attenborough is Professor in Acoustics at the Open University, a former


Editor-in-Chief of Applied Acoustics, and a former Associate Editor of the
Journal of the Acoustical Society of America and Acta Acustica. He is co-
author with Oleksandr Zaporozhets and Vadim Tokarev for Aircraft Noise
(CRC Press, 2017), and has co-authored several chapters in Environmental
Methods for Transport Noise Reduction (CRC Press, 2019). He is Chair of
ANSI S1 WG20 on the measurement of outdoor ground impedance.

Timothy Van Renterghem is Associate Professor in Environmental Sound at


Ghent University and holds a MSc. degree in Bioengineering (Environmental
Technologies) and a PhD in Applied Physical Engineering. He is Associate
Editor of Acta Acustica, the journal of the European Acoustics Association,
and Elsevier’s Urban Forestry and Urban Greening. His main research inter-
ests include the impact of local meteorology on sound propagation outdoors,
green noise reducing measures, and urban sound propagation with a strong
focus on (detailed) numerical modelling.

xvii
Chapter 1

Introduction

1.1 EARLY OBSERVATIONS

The way in which sound travels outdoors has been of interest for several cen-
turies. Initial experiments were concerned with the speed of sound [1]. In
1640, the Francisan (Minimite) friar, Marin Mersenne (1588–1648), timed
the interval between seeing the flash and hearing the report from guns fired at
a known distance and obtained a value of 450 m/s. In 1738, the French
Academy of Science used the same idea with cannon fire and reported a speed
of 332 m/s which is remarkably close to the currently accepted value for stan-
dard conditions of temperature (20°C) and pressure (at sea level) of 343 m/s.
William Derham (1657–1735), the rector of a small church near London, was
first to observe the influence of wind and temperature on sound speed and
remarked on the difference between the sound of the church bells at a certain
location over newly fallen snow compared with their sound at the same loca-
tion without snow but with a frozen ground surface.
Many records of the strange effects of the atmosphere on the propagation
of sound waves have been associated with war [2, 3]. In June 1666, Samuel
Pepys wrote that the sounds of a naval engagement between the British and
Dutch fleets were heard clearly at some spots but not at others a similar
distance away or closer. Pepys spoke to the captain of a yacht that had been
positioned between the battle and the English coast. The captain said that he
had seen the fleets and run from them, ‘…but from that hour to this hath not
heard one gun…’. The effects of the atmosphere on battle sounds were not
studied in a scientific way until after the First World War (1914–1918).
During that war, acoustic shadow zones, similar to those observed by Pepys,
were observed during the battle of Antwerp. Observers also noted that bat-
tle sounds from France only reached England during the summer months
and were best heard in Germany during the winter. After the war there was
great interest in these observations among the scientific community. Large
amounts of ammunition were detonated throughout England and the public
was asked to listen for sounds of explosions.
Although there was considerable interest in atmospheric acoustics after
the First World War, the advent of the submarine encouraged greater efforts

1
2 Predicting Outdoor Sound

in underwater acoustics research during and after the Second World War
(1939–1945). Nevertheless, subsequently, the theoretical and numerical
methods widely deployed in predicting sound propagation in the oceans
have proved to be useful in atmospheric acoustics. A meeting organized by
the University of Mississippi and held on the Mississippi Gulf Coast in 1981
was the first in which researchers in underwater acoustics met with scientists
interested in atmospheric acoustics and this has stimulated the adaptation of
the numerical methods used in underwater acoustics, for predicting sound
propagation in the atmosphere [4].

1.2 A BRIEF SURVEY OF OUTDOOR SOUND


ATTENUATION MECHANISMS

Outdoor sound is influenced by distance, by topography (including natural or


artificial barriers), by interaction with the ground and with or without vegeta-
tive ground cover and by atmospheric effects including refraction and absorp-
tion. Velocity vectors of sound and wind are additive. When the source is
downwind of the receiver, sound from the source propagates upwind. As height
in the atmosphere increases, wind speed increases and the wind speed vector
component between source and receiver is subtracted from the speed of sound
increases, leading to a negative sound speed gradient. A negative sound speed
gradient means that there is upward refraction of sound. If sound is considered
to propagate as rays, then a negative sound speed gradient causes rays to curve
upwards. When the source is near the ground during upward refraction, there
is a limiting ray that leaves the source and just grazes the ground at some point
before reaching a receiver at any given height. This defines the start of the
shadow zone. Receivers at the same distance from the source below the limiting
ray and receivers at the same height but located further away are in the shadow
zone. The distance of the start of a sound shadow from the source caused by
upward refraction depends on the sound speed gradient. However, ray tracing
(considered in more detail in Chapter 11) ceases to be valid beyond this limiting
ground-grazing ray. The shadow zone caused by upward refraction is pene-
trated by sound scattered by atmospheric turbulence thereby limiting the reduc-
tion of sound levels within the sound shadow. The many effects of atmospheric
turbulence are considered further in Chapter 10.
A negative sound speed gradient results also when the temperature
decreases with height. This is called a temperature lapse condition and is the
normal condition on a dry sunny day with little wind. A combination of
slightly negative temperature gradient, strong upwind propagation and air
absorption has been observed, in carefully monitored experiments, to reduce
sound levels, 640 m from a 6 m high source over relatively hard ground, by
up to 20 dB more than would be expected only from spherical spreading [5].
The total attenuation of a sound outdoors can be expressed as the sum of
the reductions due to geometric spreading, atmospheric absorption and the
Introduction 3

extra attenuations due to ground effects, scattering, visco-thermal effects in


vegetation, refraction in the atmosphere and diffraction by barriers.
Atmospheric absorption results from heat conduction losses, shear viscosity
losses and molecular relaxation losses and increases rapidly with frequency.
It acts as a low pass filter at long range.
Ground effects (for elevated source and receiver) are the result of interfer-
ence between sound travelling directly from source to receiver and sound
reflected from the ground. The influence of the ground depends on the source
and receiver locations as well as the nature of the ground surface. If the inter-
ference is constructive, there is an enhancement compared with the free field
level which tends to occur mainly at low frequencies. If the interference is
destructive, there is attenuation. Porous ground surfaces allow sound waves
to penetrate the pores where they are affected by viscous friction and thermal
exchanges, so the reflected sound suffers a change in phase as well as ampli-
tude. For a given source–receiver geometry, destructive interference occurs at
lower frequencies than over ground which is not porous. But, irrespective of
the porosity of the ground surface, if it is rough, there can be additional sound
attenuation due to scattering.
The root zone created by any vegetation or ground cover tends to make the
surface layer of ground more porous. Moreover, layers of partially decayed leaf
matter on the floors of forests are highly porous. Propagation through bushes,
hedges and crops involves ground effects, since they are planted on acoustically
soft ground, scattering by stems, visco-thermal scattering by foliage and, possi-
bly, acoustically induced leaf vibrations. Propagation through trees involves
reverberant scattering by tree trunks. Ground effects, scattering phenomena
and foliage attenuation are explored in more detail in Chapters 2, 5, 6 and 10.
This chapter continues by presenting data illustrating ground effects.
Subsequently, data is explored that illustrates the combined effects of ground
(including topography) and atmospheric refraction. There follow outlines of
methods for classifying and representing the variation of sound speed with
height for use in models for predicting outdoor sound propagation.

1.3 DATA ILLUSTRATING GROUND EFFECT

1.3.1 Propagation from a Fixed Jet Engine Source


Pioneering studies of the combined influences of the ground surface and mete-
orological conditions were carried out by Parkin and Scholes [6–10] using a
fixed Rolls Royce Avon jet engine as a source at two airfields (Hatfield and
Radlett). In his 1970 Rayleigh Medal Lecture, one of the investigators, the late
Peter Parkin, remarked [6],

These horizontal propagation trials showed up the ground effect, which


at first we did not believe, thinking there was something wrong with the
4 Predicting Outdoor Sound

measurements. But by listening to the jet noise at a distance, one could


clearly hear the gap in the spectrum.

These studies were among the first to quantify the change in ground effect
with type of surface. The Parkin and Scholes data showed a noticeable dif-
ference between the ground effects due to two types of grass cover. The
ground attenuation at Hatfield, although still a major propagation factor,
was less than at Radlett and its maximum value occurred at a higher fre-
quency. A change in weather conditions during their measurements also
enabled them to remark the effects of snow cover.

… measurements [were] made at Site 2 [Radlett] with 6 to 9 in. of snow on


the ground. The snow had fallen within the previous 24 hours and had not
been disturbed. The attenuations with snow on the ground were very dif-
ferent from those measured under comparable wind and temperature con-
ditions without snow … .The maximum of the ground attenuation appears
to have moved down the frequency scale by approximately 2 octaves …

Examples of the Parkin and Scholes data are shown in Figure 1.1. These
data are of the corrected level difference, i.e. the difference in sound pressure

Figure 1.1 P
 arkin and Scholes’ data for the level difference between 1.5 m high microphones
at 19 m and 347 m from a fixed jet engine source (nozzle-centre height 1.82
m) corrected for wavefront spreading and air absorption. The symbols and ⋄
represent data over airfields (grass-covered) at Radlett and Hatfield respectively
with a positive vector wind between source and receiver of 1.27 m/s (5 ft/s).
Crosses (×) represent data over approximately 0.15 m thick (6–9 in.) snow at
Hatfield with a positive vector wind of 1.52 m/s (6 ft/s).
Introduction 5

levels at 19 m (used as a reference location) and each of more distant locations


corrected for the decrease expected from spherical spreading and air absorp-
tion. They provide further evidence that ground effect is sensitive to the acousti-
cal properties of the surface which, in turn, depend on the substance of which
the surface is composed. Different ground surfaces have different porosities.
Soils have volume porosities of between 10% and 40%. Snow, which has a
porosity of around 60%, and many fibrous materials, which have porosities of
above 90%, have relatively low flow resistivities whereas a wet compacted soil
surface will have a rather high flow resistivity. Also, the thickness of the surface
porous layer is important and whether it has an acoustically hard substrate. The
Parkin and Scholes data revealed the large attenuation at low frequencies (63
Hz and 125 Hz octave bands) in the presence of thick snow. It should be noted
that, even without snow, there are significant differences in the Parkin and
Scholes’ Radlett data between summer and winter. Seasonal variations in
ground effects are discussed further in Chapter 6, section 6.4.

1.3.2 Propagation over Discontinuous Ground


The extent to which discontinuities in surface impedance can influence the
rate of attenuation is illustrated by data from measurements of noise levels
during aircraft engine run-ups made at distances of up to 3 km from the
source with the aim of defining noise contours in the vicinity of airports
[11]. Measurements were made for a range of power settings during several
summer days with near to calm weather conditions (wind speed < 5 m/s,
temperature between 20 and 25°C). Between 7 and 10 measurements were
made at every measurement station and the results were averaged. Example
jet engine noise spectra at four angles from the direction forward of the
aircraft normalized to a reference distance of 1 m are shown in Figure 1.2.

Figure 1.2 S PL Spectra,normalised to 1 m during an IL-86 engine run up test,as a function of angle
(theta) (forward of the aircraft = 0 degrees [11]. Reprinted with permission from
Elsevier.
6 Predicting Outdoor Sound

Figure 1.3 M
 easured differences (joined crosses) between the A-weighted sound level at 100
m and those measured at ranges up to 3 km during an Il-86 aircraft’s engine test
in the direction of maximum jet noise generation (~40o from exhaust axis) and
predictions for levels due to a point source at the engine centre height assuming
spherical spreading plus air absorption and various types of ground [11]. Reprinted
with permission from Elsevier.

Example measured differences in levels between a receiver at a reference


distance of 100 m and other receiver locations between 200 m and 2.7 km
are shown in Figure 1.3.
Also shown in Figure 1.3 are predictions based on models detailed in
Chapters 2 and 5. The predictions use equation (2.40) and a one-parameter
impedance model (see Chapter 5, equations (5.1) and (5.2)), assuming effec-
tive flow resistivities of 20,000 kPa s m–2, 300 kPa s m–2 and 2000 kPa s m–2,
respectively, for concrete, grass and soil. The predictions show that, up to
500 m distance, the data are consistent with propagation in acoustically
neutral conditions over an acoustically hard surface but beyond 1 km are
more representative of levels predicted over an acoustically soft surface.
Figure 1.4 shows similar data and predictions for the attenuation from a
propeller aircraft.
The averaged data in both the maximum jet noise and the maximum pro-
peller noise directions indicate a change from one rate of attenuation to
another at distances greater than 500 m which can be attributed to a varia-
tion with range in the nature of the ground surface. The run-ups took place
over the concrete surface of an apron. Further away (i.e. between 500 m and
700 m from the aircraft in various directions), the ground surface was ‘soil’
and/or ‘grass’. A good fit to the data is obtained by predictions that include
an impedance discontinuity between 500 and 1000 m from the source [11].
Introduction 7

Figure 1.4 M
 easured differences (joined crosses) between the A-weighted sound level
at 100m and those measured at ranges up to 3km from a turbo-prop engine
(on an An-24 aircraft) in the direction of maximum propeller noise generation
(~80o from axis of engine inlet) [11]. Reprinted with permission from Elsevier.

Figure 1.5 M
 easured sound levels in the direction of maximum propeller noise from a
turboprop aircraft and the fit given by equation (1.1).

The empirical fit shown in Figure 1.5 to the data beyond 100 m in the
direction of maximum propeller noise is given by

LA  162  27.5 log  d   0.005d, (1.1)

where d (m) is the horizontal distance.


8 Predicting Outdoor Sound

1.4 DATA ILLUSTRATING THE COMBINED EFFECTS


OF GROUND AND METEOROLOGY

1.4.1 More Fixed Jet Engine Data


Although the wind speed was measured, the classical experiments by Parkin
and Scholes involved relatively little meteorological monitoring. The impor-
tant role of atmospheric turbulence was not appreciated at the time, so, for
example, the fine-scale fluctuations in wind speed were not monitored. Similar
measurements to those carried out by Parkin and Scholes using a fixed jet
engine source have been made but they were augmented by more comprehen-
sive meteorological data. The fixed jet engine was operated as a broadband
noise source at a Rolls Royce test facility in a disused airfield at Hucknall,
Nottingham, UK [12]. Simultaneous acoustic and meteorological measurements
were made. In addition to wind and temperature gradient measurements, the
fluctuation in wind velocity measurements was recorded and used as a mea-
sure of turbulence. Some of the data obtained under low wind and low turbu-
lence conditions over continuous grassland are shown in Figure 1.6. Also
shown is the third octave power spectrum of the Avon engine source between
100 Hz and 4000 Hz deduced from the measured spectrum at 152.4 m after
correcting for spherical spreading and ground effect. The data obtained at the
longest range is limited by background noise above 3 kHz. The significant
dips in the received spectra between 100 and 500 Hz are clear evidence of
ground effect. However, it is noticeable that the ground effect at Hucknall is
different from that measured at either Radlett or Hatfield.
The influence of small changes in the wind speed and turbulence strength
on the measured spectra at the longest range is demonstrated in Figure 1.7.
The associated meteorological conditions are detailed in Table 1.1. The
ground effect between 100 Hz and 400 Hz is relatively stable and signifi-
cantly greater at the low microphones where it is shifted in frequency com-
pared with ground effect at the high microphones. The data for both
microphone heights show considerable variability between 400 Hz and 2
kHz which can be attributed to changes in wind velocity and turbulence.
Figures 1.8 and 1.9 show A-weighted levels deduced from average spectra
in consecutive 26 s periods measured at 1.2 m height and ranges of 152.4 m,
457.6 m, 762.2 m and 1158.4 m over grassland at Hucknall. Figure 1.8 shows
data for low wind speed (less than 2 m/s from source to receiver) and low
turbulence conditions. Figure 1.9 shows data for moderate downwind condi-
tions (approximately 6 m/s from source to receiver) and for higher turbulence
intensities. The details of the meteorological conditions corresponding to
Figures 1.8 and 1.9 are listed in Tables 1.2 and 1.3, respectively.
Figure 1.8 shows the considerable spread in the measured levels at the
longer ranges resulting from the variation in wind speed and direction (up
to approximately 2 m/s downwind at 6.4 m height) and turbulence levels.
The data for stronger downwind conditions (up to approximately 6.5 m/s at
6.4 m height) in Figure 1.9 indicate consistently higher levels than those dur-
ing the relatively low wind speed conditions. On the other hand, they have
Introduction 9

Figure 1.6 D
 ata recorded at 1.2 m high receivers at horizontal ranges of 152.4 m (solid
line), 457 m (dotted line), 762 m (dashed line) and 1158 m (dash-dot line) from
a fixed Rolls Royce jet engine source with the nozzle centre 2.16 m above
an airfield at Hucknall, Notts. These data represent simultaneous recordings
averaged over 26 s during zero wind and low turbulence conditions (block 20
of run 454, see Figure 1.3). Also shown (connected circles) is the deduced third
octave power spectrum of the Avon jet engine source after subtracting 50 dB.

a smaller spread. Although only four averages are shown in Figure 1.9, their
spread is smaller than for any four averages exhibited in Figure 1.8. This is
consistent with the assertion in ISO 9613-2 [13] that the variation in sound
levels is less under ‘moderate’ downwind conditions. The average down-
wind level measured at Hucknall is about 10 dB higher than the levels for
the lowest wind speed and turbulence conditions at 1.1 km from the source.

1.4.2 Road Traffic Noise Propagation over Flat


Terrain under Strong Temperature Inversion
Figure 1.10 shows temperature profiles derived from series of air temperature
measurements at different heights (up to 13 m) near Scottsdale in the Phoenix
valley (Arizona, US) [13]. The area can be characterized meteorologically as
one with very light synoptic winds and clear skies, giving rise to strong diurnal
10 Predicting Outdoor Sound

Figure 1.7 Simultaneously measured narrow band (25 Hz interval) spectra at low
(1.2 m – upper graph) and high (6.4 m – lower graph) microphones between
50 Hz and 10 kHz at 1158.2 m from a fixed Avon jet engine source averaged
over 26 s intervals during low wind, low turbulence conditions at Hucknall
(Notts. UK). The conditions are specified in Table 1.1 and the key.
Another random document with
no related content on Scribd:
The Project Gutenberg eBook of Refraction and
muscular imbalance, as simplified through the
use of the ski-optometer
This ebook is for the use of anyone anywhere in the United States
and most other parts of the world at no cost and with almost no
restrictions whatsoever. You may copy it, give it away or re-use it
under the terms of the Project Gutenberg License included with this
ebook or online at www.gutenberg.org. If you are not located in the
United States, you will have to check the laws of the country where
you are located before using this eBook.

Title: Refraction and muscular imbalance, as simplified through the


use of the ski-optometer

Author: Daniel Woolf

Release date: August 29, 2023 [eBook #71517]

Language: English

Original publication: New York: Theodore S. Holbrook, 1921

Credits: deaurider and the Online Distributed Proofreading Team at


https://www.pgdp.net (This file was produced from images
generously made available by The Internet Archive)

*** START OF THE PROJECT GUTENBERG EBOOK


REFRACTION AND MUSCULAR IMBALANCE, AS SIMPLIFIED
THROUGH THE USE OF THE SKI-OPTOMETER ***
Ski-optometer Master Model 215
Embodying in a Single Instrument, in Convenient Form,
Cylindrical and Spherical Lenses, in Combination
with Appliances for Testing and Correcting
Muscular Imbalance.

Refraction and
Muscular Imbalance
As Simplified Through the Use
of the Ski-optometer

By

DANIEL WOOLF
WOOLF INSTRUMENT CORPORATION
New York: 516 Fifth Avenue

Copyright 1921
By WOOLF INSTRUMENT CORPORATION

Published by
Theodore S. Holbrook
New York
CONTENTS
Page
Chapter I
Ski-optometer Construction 1
Convex Spherical Lenses 2
Operates and Indicates Automatically 6
Concave Spherical Lenses 7

Chapter II
Cylindrical Lenses 10
Obtaining Correct Focus 11
Why Concave Cylinders Are Used Exclusively 14
Transposition of Lenses 14

Chapter III
How the Ski-optometer Assists in Refraction 17
The Use of the Ski-optometer in Skioscopy 17
A Simplified Skioscopic Method 20
Employing Spheres and Cylinders in Skioscopy 22
Use of the Ski-optometer in Subjective Testing 23
A Simplified Subjective Method 24
Procedure for Using Minus Cylinders Exclusively 26
Constant Attention Not Required 29

Chapter IV
Important Points in Connection with the
Use of the Ski-optometer 30
Elimination of Trial-Frame Discomfort 30
Rigidity of Construction 31
How to Place the Ski-optometer in Position 32
Cleaning the Lenses 33
Accuracy Assured in Every Test 34
Built to Last a Lifetime 35
Chapter V
Condensed Procedure for Making Sphere and
Cylinder Test with the Ski-optometer 37
Subjective Distance Test 37
Subjective Reading Test 40

Chapter VI
Muscular Imbalance 41
The Action of Prisms 42
The Phorometer 43
The Maddox Rod 44
Procedure for Making the Muscle Test 45
Binocular and Monocular Test 47

Chapter VII
The Binocular Muscle Test 48
Made with the Maddox Rod and Phorometer 48
Esophoria and Exophoria 50
Making Muscle Test Before and After Optical Correction 52
When to Consider Correction of Muscular Imbalance 53
Four Methods for Correction of Muscular Imbalance 54
The Rotary Prism 54
Use of the Rotary Prism in Binocular Muscle Tests 56

Chapter VIII
The Monocular Duction Muscle Test 58
Made with Both Rotary Prisms 58
Locating the Faulty Muscle 58
Adduction 59
Abduction 61
Superduction 62
Subduction 63
Procedure for Monocular Muscle Testing 64
Diagnosing a Specific Muscle Case 65
Chapter IX
First Method of Treatment—Optical Correction 70
Esophoria 70
Treatment for Correcting Esophoria in Children 72
How Optical Correction Tends to Decrease 6°
Esophoria in a Child 74

Chapter X
Second Method of Treatment—Muscular Exercise 75
Made with Two Rotary Prisms and Red Maddox Rod 75
Exophoria 75
An Assumed Case 78
Effect of Muscular Exercise 80
Home Treatment for Muscular Exercise—
Square Prism Set Used in Conjunction with
the Ski-optometer 82

Chapter XI
Third Method of Treatment—Prism Lenses 84
When and How Employed 84
Prism Reduction Method 85

Chapter XII
A Condensation of Previous Chapters on the Procedure
for Muscle Testing with the Ski-optometer 87
Four Methods of Treating an Imbalance Case when
the Preceding One Fails 90
Prisms 92
Cyclophoria 92

Chapter XIII
Cyclophoria 93
Made with Maddox Rods and Rotary Prisms 93

Chapter XIV
Cycloduction Test 99
Made with the Combined Use of the Two Maddox Rods 99
Treatment for Cyclophoria 102

Chapter XV
Movements of the Eyeballs and their Anomalies 105
Monocular Fixation 105
Binocular Fixation 106
Orthophoria 107
Heterophoria 107
Squint 108
Varieties of Heterophoria and Squint 109

Chapter XVI
Law of Projection 114
Suppression of Image 115
Monocular Diplopia 115
Table of Diplopia 116
Movement of Each Eye Singly 117
Subsidiary Actions 118
Field of Action of Muscles 120
Direction of the Gaze 120
Primary Position—Field of Fixation 121
Binocular Movements 121
Parallel Movements 122
Lateral Rotators 123
Eye Associates 124
Movements of Convergence 125
Movements of Divergence 125
Vertical Divergence 126
Orthophoria 126
Heterophoria 126
Subdivisions 126

Chapter XVII
Symptoms of Heterophoria 128
Treatment 130
Destrophoria and Laevophoria 132
The demands of the day for maximum efficiency in
the refracting world are largely accountable for the
inception, continuous improvement and ultimate
development of the master model Ski-optometer.
The present volume, dealing with the instrument’s
distinctive operative features, has been prepared not
only for Ski-optometer users, but also for those
interested in the simplification of refraction and
muscular imbalance.
The author is indebted for invaluable counsel, to

Louis J. Ameno, M.D., New York.


E. LeRoy Ryer, O.D., New York.
Jos. D. Heitger, M.D., Louisville, Ky.
W. B. Needles, N.D., Kansas City, Mo.
INTRODUCTORY

W
hile in a measure the conventional trial-case still serves its
purpose, so much of the refractionist’s time is consumed
through the mechanical process of individually transferring the
trial-case spheres and cylinder lenses, that far too little thought is
given to muscular imbalance, notwithstanding its importance in all
refraction cases.
Dr. Samuel Theibold, of Johns Hopkins University, in a recent
address before the American Medical Association, stated that the
average refractionist was inclined to devote an excess of time to
general refraction, completely overlooking the important test and
correction of muscular imbalance. If the latter is to be at all
considered, general refraction must be simplified—without impairing
its accuracy—a result that is greatly facilitated through the use of the
Ski-optometer.
One must admit that tediously selecting the required trial-case
lens—whether sphere, cylinder or prism—watching the stamped
number on the handle—continual wiping and inserting each
individual lens in a trial-frame is a time-consuming practise. This is
readily overcome, however, through the employment of the Ski-
optometer.
In a word, the Ski-optometer is practically an automatic trial-case,
bearing the same relation to the refracting room as the accepted
labor and time-saving devices of the day bear to the commercial
world.
The present volume has accordingly been published, not alone in
the interest of those possessing a Ski-optometer, but also for those
interested in attaining the highest point of efficiency in the work of
refraction and muscular imbalance.
Ski-optometer Lens Battery (almost actual size)
showing how sphere and cylinder lenses are
procured.
After obtaining FINAL results, your prescription is
automatically registered,
ALL READY for you to transcribe.
Fig. 1—The three time-saving moves necessary in
the operation of the Ski-optometer.
Chapter I
SKI-OPTOMETER CONSTRUCTION

A
far better understanding of the instrument will be secured if the
refractionist possessing a Ski-optometer will place it before him,
working out each operation and experiment step by step in its
proper routine.
The three moves as outlined in Fig. 1 should first be thoughtfully
studied and the method of obtaining the spheres and cylinders
carefully observed.
Fig. 2—To Obtain Plano.
1—Set spherical indicator at “000” as illustrated above.
2—Set cylinder indicator to “0”.
3—Set pointer of supplementary disk at “open”.
The instrument should then be set at zero or “plano,” a position
indicated by the appearance of the three “0 0 0” at the spherical
register, in conjunction with one “0” or zero, for the cylinder at its
register, marked “CC Cyl.”
After this move, the supplementary disk’s pointer should be set at
“open” (Fig. 2).
Fig. 3—To obtain sphericals, turn this
Single Reel as shown by dotted finger. This
assures an automatic and simultaneous
registration at sphere indicator of focus of
lens appearing at sight opening.

Convex Spherical Lenses


A careful study will show that the Ski-optometer’s spherical lenses
are obtained by merely turning the smaller reel (Fig. 3). The first
outward turn of this reel, toward the temporal side of the instrument,
draws into position in regular order the spherical lenses +.25, +.50,
+.75, and +1.D., as shown in Fig. 3a.
3-A—Outer spherical reel containing Cx. sphericals
from .025 to 1.00D and a blank.
3-B—Inner spherical disk containing Cx. sphericals,
automatically turns within 3-A.

3-C—Supplementary spherical disk.


By means of a concealed tooth gear, an inner disk is automatically
picked up, placing its first lens +1.25D in position (Fig. 3b). This
+1.25D spherical lens remains stationary while the outer disk again
revolves, adding to it the original +.25, +.50, +.75 and +1.D., the latter
totalling +2.25D. At this point, the instrument again automatically picks
up its inner disk, thereby placing its second lens, +2.50D, in position.

You might also like