You are on page 1of 107

COACHING SWIMMING

AN INTRODUCTORY MANUAL
Second Edition

TEXTBOOK FOR ‘BRONZE COACHING LICENCE’


NATIONAL COACHING ACCREDIATION COURSE

Dr Ralph J. Richards

Copyright © 2004 by the Australian Swimming Coaches and Teachers Association (ASCTA) All
rights reserved. No portion of this book may be reproduced in any form without written
permission from the publisher.

ISBN 0 000 00000 0


Front Cover photos taken by Ralph Richards

ii

Acknowledgement

Coaching education in Australia has come a long way during the past 25 years. The
Australian system of educating, training, and accrediting coaches is now highly regarded
worldwide. Two key factors have made the Australian system successful. First,
developing new coaches is done within a practical framework, the emphasis is clearly on
applied principles. Second, coaches are stakeholders in the accreditation process;
Australian Swimming Inc. in cooperation with the Australian Swimming Coaches and
Teachers Association work closely to deliver training and continuing education. Both
organisations encourage swimming coaches to become more professional in their
approach. Good coaching provides the day-to-day drive that keeps our sport moving
forward. Successful coaches must exercise many skills; including a sharp eye for good
swimming technique, an inquiring mind that applies scientific principles, and great
interpersonal skills to work effectively with athletes.

This textbook will assist beginning coaches in taking one more step along the pathway of
personal improvement. It offers another look at some of the basic concepts introduced
during the entry-level coach qualification course. The emphasis at the Bronze Licence
level begins to take the coach in the pursuit of more advanced swimming performance.
This text is not a comprehensive collection of coaching information, there is still much
more to learn and apply. Most coaches will continue to refine their coaching knowledge
and skills through ongoing study and practical learning experiences once their
accreditation status has been confirmed. One fundamental way to supplement the
information gathered during a course is to observe other coaches and then discuss their
coaching methods and ideas. I have worked with many coaches during my career and
each has contributed something to my style and understanding of coaching.

It would be impossible to list everyone having an influence on my professional


development as a coach, yet remiss of me not to mention at least two coaches. First, the
legendary swimming coach, James ‘Doc’ Councilman, has been the greatest influence on
my thinking and inspiration for me to pursue an understanding of science in swimming. I
had the pleasure of working under Doc for several years and his influence remains with
me to this day. The second significant coach is a contemporary of mine, Bill
Sweetenham. His passionate pursuit of excellence has served as an example to me, as
well as a generation of his peers. In addition, there are many past and current Australian
coaches, sports scientists, and administrators who have helped me to develop my
coaching skills; I offer a sincere ‘thank you’ for your contribution.

Australian Swimming Inc. and the Australian Swimming Coaches and Teachers
Association are united in their belief that well prepared, educated, ethical, and motivated
coaches are the cornerstone of Australia’s swimming success. More importantly,
dedicated coaches will continue to encourage and inspire athletes to become great persons
in their own right, regardless of their swimming prowess. Every coach can take great
satisfaction in their work as they strive to be the best they can be. I encourage you to
embrace your coaching experience and pursue it with enthusiasm. Good Luck. Dr Ralph
J. Richards
iii

Foreword

It has been said that there are three ‘T’s to successful swimming: Technique – Technique
– Technique. However, successful swimming coaching is about five key ‘I’s: Information
– Inspiration – Intuition – Instinct – and Imagination.

In this book, Dr Richards presents the information relevant to the coaching of swimming
better than almost anyone else can. Ralph’s vast knowledge of the sport of swimming, his
understanding of swimming technique and the skills of successful swimmers, his
experience, and his passion for swimming has made him one of Australia’s finest
swimming coaches and a world leading educator of swimming coaches. The information
presented is the best there is: current, logical, accurate, comprehensive, detailed – it is
outstanding.

As a coach you should be inspired to take this information and use it to help swimmers
achieve their goals to the extent of their talents and to the limits of their dreams.

You should use your instincts and intuition to apply this knowledge effectively to each
individual (the sixth ‘I’) swimmer. Coaching is a people game; it’s about working with
people and for people. Coaching is part intellect, but mostly feeling; it’s about living the
dream of the swimmers working with you.

And coaching is about imagination – imagination is the element that will drive our sport
forward and take it to places we have only dreamed of.

Many people will read this book. Most will take away something that will make them
more knowledgeable about swimming, about coaching, about sports science, or some
other aspect of aquatics. Some will use the knowledge gained from this book; together
with their own inspiration, intuition, instinct and imagination; to achieve remarkable
things as a coach. Which one will you be?

Ralph’s expertise is recognised internationally and his impact on the sport of swimming
in Australia; through his coaching, teaching, mentoring, writing and lecturing; has been
considerable. As a Club coach, an Elite coach, a National Team coach, a Teacher, and a
Leader he has achieved remarkable success over the past thirty years. However, I’m
certain his most lasting impact is yet to come – not the least being the coaching outcomes
achieved by those implementing the five ‘I’s.
Wayne Goldsmith

iv

Contents
Acknowledgment iii Forward iv

1 Improving Coaching Skills 1


Key Concepts Transition from Teaching
to Coaching Phases of Swimmer
Development Adaptation Coaching
Junior Squads Developmental
Considerations Age-Group Training
Model Summary of Key Points

2 Mechanics of Swimming 10
General Terminology Buoyancy and
Weight Inertia and Momentum Levers
and the Kinetic Chain Velocity and
Acceleration Continuity of
Momentum Propulsive Forces
Resistance Forces Explaining
Swimming Movements
3 Freestyle and Backstroke 18
Freestyle:
Overview Kick Armstroke –
Propulsive Phase Armstroke –
Recovery Phase Breathing Timing and
Rhythm Backstroke:
Overview Kick Armstroke –
Propulsive Phase Armstroke –
Recovery Phase Breathing

4 Breaststroke and Butterfly 26


Breaststroke:
Overview Streamlining
Armstroke – Propulsive Phase

v
Arm Recovery and Breathing Kick
Timing Butterfly:
Overview Streamlining Kick
Armstroke – Propulsive Phase
Armstroke – Recovery Phase
Breathing Timing

5 Starts, Turns, and Finishes 34


Dive Start
Dive Start Checklist Backstroke Start
Backstroke Start Checklist Racing Turns
(Overview) Freestyle Tumble Turn
Freestyle Turn Checklist Backstroke Tumble
Turn
Backstroke Turn Checklist Breaststroke Turn
Breaststroke Turn Checklist Butterfly Turn
Butterfly Turn Checklist Individual Medley
Turns Race Finishes
Finishing Checklist

6 Physiological Considerations 46
Energy Supply for Swimming
Interaction of the Energy Systems
Energy System Capacities
Cardiovascular Considerations
Measuring Exercise Intensity Fatigue

7 Training Methods 56
Aerobic Base Training Aerobic
Endurance Training Critical
Velocity Training Maximum
Aerobic Training Lactate
Tolerance Training Peak
Lactate Training Sprint
Training Maximum Speed
Training

vi

8 Drills and Stroke-Rate 71


Stroke Development Drills Stroke
Correction Drills Linking Drills Speed
Drills Adding Value and Variety to
Drills Stroke-Rate

9 Long Term Swimmer Development 80


Identifying Talented Swimmers Break Point Volume
Concept Multi-Year Age-Group Swimmer Development
Model

10 Strength and Flexibility 85


Considerations for Young Swimmers
Strength Training Outcomes Male and
Female Differences Muscle Loading
Relationships Periodisation of Strength
Training Exercise Selection Stretching

11 Planning Training 94
Season Outline Training Phases
Weekly Training Cycles
Individual Training Sessions
Training Objectives Examples of
Session Plans Early Season
Considerations Mid-Season
Considerations Late Season
Considerations Transition
Between Seasons Summary The
Integrated Training Model

12 Mental Skills 111


Coaching Style
Communication
Applied Approach
Psychological Skills
Strategies

13 Health and Injury Prevention 116


The Overtrained State
Recovery Process

vii
Health Considerations
Injury Considerations

14 Nutrition 123
Objectives Everyday Diet Tips on
Food Selection Fluid and Glycogen
Replenishment Tips on Maintaining
Body Weight Vitamins and Minerals
Competition Day Alcohol

15 Drugs in Sport 128


Overview Banned Substances and
Methods Drug Testing Why are
Substances Banned? Drugs
Education

References 132
viii
1

CHAPTER 1 IMPROVING
COACHING SKILLS

The roles and responsibilities of a swimming coach today are certainly different from 10,
20 or 30 years ago. Coaching, whether as a vocation or on a volunteer basis, has become
much more "professional" in terms of our application of knowledge and skills. Although
many factors contribute to swimming performance, coaching effectiveness is certainly a
major influence. Today’s club level swimmer probably exceeds the performance standard
of state level swimmers from a generation ago. Environmental, social, and genetic factors
alone could never account for the sustained rate of improvement within our sport.
Therefore, the influence of the coach, since he/she has direct control over training design
and implementation, must be a key factor.

There are many challenges facing the coach, first and foremost is the ability to
communicate ideas so that efficient swimming skills are developed at a young age. The
Green Licence coach will demonstrate skills that allow him/her to effectively
communicate with swimmers to teach skills and organise basic training activities. This is
sometimes called the “art” of coaching. A Bronze Licence coach must demonstrate
greater depth of knowledge across a wider range of factors that influence swimming
performance. This will require a theoretical base that embraces scientific principles and
builds upon practical experience.

Key Concepts

This textbook will reinforce four key concepts that impact upon coaching junior
swimmers: (1) the principle of adaptation, (2) mechanical principles that shape propulsive
movement in the water, (3) mental skills that may enhance performance, and (4)
maturational considerations. Experienced coaches know that the best training facilities,
the best training plans, and the greatest natural physical ability do not guarantee success;
they only improve the chances of being successful. Human behaviour patterns influence
everything the swimmer and the coach do or plan to do. This is why the Green Licence
course placed such a heavy emphasis on understanding an athlete’s behaviour and
developing good communication skills to shape that behaviour. Above everything else the
swimmer and coach must have a shared understanding of what they are trying to achieve.

Transition from Teaching to Coaching

The initial focus of aquatic instruction must always be safety and learning to enjoy water
activities. Aquatic instruction then progresses through various stages where the swimmer
gains confidence and independence of movement in the water. Consistent skill in
performing the basic swimming strokes, usually backstroke and freestyle, begins to
emerge. The swimmer then acquires greater diversity in movement skill by mastering
breaststroke and then butterfly. An accomplished swimmer usually emerges from this
learning progression between the ages of 8 and 11 years of age.
Chapter 1 Improving Coaching Skills Coaching Swimming: An
Introductory Manual (second edition)

2 During the advanced stages of this learn-to-swim progression the departure point from a
purely learning activity to a performance-based activity is reached. This defining point is
not a discrete event, but a gradual process of introducing greater emphasis on
massed-practice and performance outcomes. The question of ‘how much’ swimming can
and should be done at each stage of development is central to good planning. Learning
new skills and consolidating these skills into established motor patterns requires repeated
practice. During a lesson (or training session) of 30 – 45 – or 60 minutes duration there is
a gradual increase in the total swimming distance completed. Sessions are also defined by
the number of skills utilized and the quality of their execution. There are also purely ‘fun’
activities that must be included in the mix. In this context ‘work’ is defined by the
physiological requirements of energy production that produces movement and is
generally quantified in terms of the number of metres swum and the relative intensity at
which that swimming is performed.

Phases of Swimmer Development

When a swimmer completes lean-to-swim lessons he/she should have the basic technical
skill to swim competitively. Development as a swimming athlete is determined by the
combination of training and competition opportunities presented and one’s individual
talent. We might view a swimmer’s career progress in terms of three phases of
development.

First, the swimmer must learn to train; this is called the ‘training phase’. This phase will
last several years and will take the swimmer through the pubertal growth spurt and
biological maturation. Competition is also an important part of the swimmer’s overall
development, but training progressions are the key element at this stage of a swimmer’s
career. Many talented swimmers do not distinguish themselves in competition during
this phase because they may be late maturing. The ideal training background includes a
good work ethic, positive attitude, love of the sport, and determination to succeed.

Second, the swimmer goes through a phase where competitive results begin to identify
those individuals having the potential to rise to elite level. This does not mean that a
swimmer who is not winning age-group medals will never reach the top of the sport.
However, the likelihood that well preformed age-group swimmers will advance to senior
swimming is established during this phase.

Third, the final phase of a swimmer’s career is the ‘performance phase’. Only a small
percentage of competitive swimmers advance to this phase of mature sporting
development. Most swimmers remain at the second phase, even those who advance
through age-group ranks and compete at open level. Competitive swimmers can realise
all the benefits (i.e. social, skill development, and fitness) from sport without
progressing to the final stage of elite development.

Chapter 1 Improving Coaching Skills Coaching Swimming: An


Introductory Manual (second edition)
3
Three Phases of Swimmer Development – Training Outcomes
TRAINING COMPONENTS:
1. TRAINING
PHASE
2. COMPETITION
PHASE
3. PERFORMANCE
PHASE Overall Training Objective
Learning the Basic Training Fundamentals
Progressive Build-up (volume then intensity)
Systematic High Level Training Associated Stage of Physical Maturity
Prepubescent and Early Puberty
Pubescent and Post- Pubescent
Physically Mature Athlete
Movement Skill Development
Refine Fundamental Movement Skills
Master All Skills At Race Pace/Pressure
Consistent Skill App- lication at All Times Technical Model Acquire Basic Skills in
all Four Strokes
Advanced Skill Level in all Four Strokes
Maintain General Skills and Specialise Knowledge “How To” Swim Strokes,
Starts, Turns
Race Tactics & Pacing Strategies
Performance Analysis Strengths & Weakness Sportsmanship Respect for Team-mates
and Coaches
Respect for other Com- petitors & Officials
Interaction with Sport (media, public, etc.) Personal Interactions Work with Coach and
Team-mates
Support Club & Team Goals
Demonstrate Leadership
Nutrition Understand Principles of
Good Nutrition
Use Best Practice and Monitor Eating Habits
Maintain Body Weight and Health Aerobic Conditioning Acquire Training Back-
ground
Increase Volume and Intensity of Training
Maintain Aerobic Fitness
Anaerobic Conditioning Maintain Stroke Tech-
nique during Sprints
Increase Volume of Intense Training
Increase Quality of Intense Training Muscular Strength Core Body Strength and
Muscle Control
Increase Strength -- Balanced Development
Develop Specific Strength & Power Flexibility and Range of Movement
Learn Swim Technique with Range of Motion
Maintain/Increase Joint Flexibility & Stability
Maintain all Elements of Muscle/Joint Action Recovery and Regeneration
Understand the Role of Recovery
Apply a Variety of Recovery Methods
Develop Individual Recovery Routine Emotional & Psycho- logical Development
Enjoy Swimming Experiences
Control Mental State during Competition
Develop Mental Skills to Meet any Situation Medical Control (monitoring)
Check Growth, Posture & Body Structure
Apply Injury Prevention Strategies
Monitor Health Status & Use Rehab Techniques [adapted from “A Plan Behind the Dream”, Vern Gambetta, ASCTA Journal,
Vol. 15, No. 1]
Adaptation
Adaptation is a process of change which seeks to meet the demands placed upon an individual
from both physical and psychological sources. The process begins when a stimulus is introduced
which forces the individual to counter with a response. In the simplest of circumstances if the
response is sufficient there is no need for adaptation. However, in real situations even the
smallest task is a collection of thousands of stimuli that must be interpreted and acted upon.
Repeated exposure to stimuli produces a net result that strengthens with time; this is the process
of adaptation. Coaches use this broad concept to manipulate stimuli so that some of them will be
beyond the immediate response capability of the individual. If the loading (i.e. the volume and
rate of introduced stimuli) is progressively increased, over time the individual will change or
adapt so that a greater number of suitable responses are possible. A key element of this process is
the recovery or ‘unloading’ period which follows the stimulus-response to complete the
sequence. It's during the recovery period that adaptation actually takes place.
When coaches deal with complete tasks, such as training a swimmer for a specific event, the
progressive overload concept is applied. For example, if a swimmer has a current capability of
swimming 50m in 40sec, but wants to swim that distance in 35sec, what must be done? The
swimmer performs a number of repeated swims,
Chapter 1 Improving Coaching Skills Coaching Swimming: An Introductory Manual (second edition)

4 which may or may not be at the target distance, the total stimulus will be greater than
the requirement for a single 40sec 50m effort. During the recovery period the body
changes in many ways as a direct result of the total stimulus and, provided the body can
adjust, over time the capability will improve to 39sec, then 38sec, etc., until the target is
achieved. It must be remembered that each time a new 'capability' is achieved, the total
loading must also be reset to a higher level; otherwise adaptation does not continue.
Naturally, this is a very simple example used to illustrate a complex process which
involves every facet of performance. Every swimmer is constantly being exposed to an
overwhelming number of physical and psychological stimuli, some of them contributing
to sporting adaptation and some of them detracting from it.

The concept of specificity suggests that when a greater proportion of stimuli that make up
the total loading are directly related to producing the desired adaptation, the resulting
adaptation will take place more quickly and completely. In the above example, we can
improve the swimmer's capacity to race at 50m by training repeat 100m or 200m
distances. This will positively influence a number of mechanisms involving energy
production and muscle endurance to allow performance improvement. However, a
different training emphasis, for example repeat 25m swims, will also produce the desired
performance changes (i.e. improvement of 50m swim time), but utilise performance
mechanisms specific to the requirements of the event. Each approach produces some of
the desired performance outcomes, but neither completely satisfies all training
requirements because rarely do training outcomes have such a narrow focus. The first
training method will not stimulate enough pure speed. The second training method will
not adequately stimulate a range of physiological improvements that will prepare the
swimmer for future adaptations to improve speed. Only by combining the two training
methods (which may seem the antithesis of specificity) do we achieve the best possible
outcome. Specific speed improvements now, and continued improvement into the future.
As higher performance levels are reached (i.e. elite), the concept of specificity takes on
additional considerations to focus training outcomes more precisely to the demands of an
event. However, elite performance is always the culmination of many specific factors that
interact over a long period of time.

The final concept that's of general importance is periodisation. This term refers to the
sequencing of events to produce maximum adaptation. The rate of adaptation can be
accelerated by using a number of techniques. First, if the stimuli are constantly changing
they will challenge the body to adapt. This does not mean the introduction of stimuli
designed to produce different types of adaptation, but stimuli that are varied or slightly
different, and yet producing the same result. Second, periodisation is used to structure the
logical development of physical and mental capacities. To use a simple analogy; the
foundation of a building must precede the framework, which precedes the finished walls.
Third, the concept of periodisation is applied by the coach to plan, monitor, and assess the
success of the training program over time; within a season and over successive seasons.
Modern training programs are multi- dimensional; that is, they incorporate many training
outcomes from many different physiological and psychological sources.

If there is one lesson a swim coach must learn, it's that nothing is ever as simple as it
seems; one's depth of understanding helps to make complex programming

Chapter 1 Improving Coaching Skills Coaching Swimming: An


Introductory Manual (second edition)
5

manageable. It's the coach's ability for personal growth, including a better understanding
of coaching methodology from year to year, that serves as the basis for adaptation to
higher and higher levels of coaching expertise.

Coaching Junior Squads

Most parents have several basic concerns regarding their child’s involvement in a
swimming training program. What is the focus of the training program – fitness, skill, or
competition? How do training commitments change with age and performance? How
many sessions are appropriate each week and what is the content of those sessions (i.e.
volume and intensity)? What is an appropriate competition program? Whilst there is no
single right or wrong answer to each question, there are a number of supporting
recommendations that should be taken into account.

Prior to puberty the paramount objectives of any program should be to instil a love of the
sport, teach quality technique, and develop all-around skill. Junior training programs must
be well thought-out and have goals that include high-level skill development and
appropriate emphasis on physiological conditioning. Decisions regarding volume and
intensity of training and the importance of competitions during early childhood involve
complex issues. Either over-exposure or under-exposure to training may detract from
achieving one's full potential. Age-group competitive swimming must always be seen as a
means to an end, not an end to itself. Finding the right answers usually comes down to
what is suitable for each individual within the general framework of developmental
principles.

The composition of major Australian Swimming Teams (i.e. Olympics, World


Championships, Commonwealth Games, Pan-Pac Championships) in recent years has
reflected the trend toward older, mature competitors, staying in our sport. This highlights
the need for coaches of junior squads to be aware of their role during the early stages of
development and the transition that takes place when physical maturity and a growing
interest in performance swimming emerges. Peak performance at international level is the
result of many things; including early skill development (physical, technical and
psychological skills), maturity, and the optimum development of physiological capacities.

Training is an ill-defined term during a swimmer's early years of development because


the acquisition and refinement of swimming skill is still ongoing. However, training
involves regular participation in a program that has at least three broad-based
performance objectives. The first and foremost objective will always be the continued
acquisition and refinement of movement skills that contribute to swimming efficiency.
The second objective is an increase in the physiological capacities that will allow faster
swimming speed and/or greater endurance. The third objective is to create a positive
psychological experience for every swimmer. This provides the internal motivation for a
swimmer to dedicate the time and effort required for continual improvement. If all three
objectives are achieved, the swimmer is more likely to remain on a track of positive
improvement. The individual is able to swim further, faster, more skilfully, and cope with
the demands of training and competition.

Chapter 1 Improving Coaching Skills Coaching Swimming: An


Introductory Manual (second edition)

6 Coaches working with junior squads should have an overall understanding of what
swimmers can realistically achieve. Initially this understanding comes from knowledge of
appropriate coaching methodology and maturational considerations. As the coach gains
more practical experience the accumulation of empirical knowledge will shape the
decision making process. Appropriate strategies are then implemented to achieve the
three performance objectives. Research has highlighted the fact that young children place
great value upon learning skills and active participation (i.e. the element of 'doing') much
more than receiving external rewards. The top five factors identified by children as
influencing their continued participation in sport are: 1. Learning new skills and
improving existing skills. 2. Using their skills to improve performance. 3. Having an
enjoyable experience with the coach/instructor. 4. Testing their ability by competing with
their peers. 5. Being with friends and social interaction.

If we accept that skill development is essential, the coach must develop a good sense of
what is acceptable swimming mechanics, the ‘educated eye’ for correct technique. This
also means that a coach working with young swimmers must have appropriate
communication skills to convey the desired messages. The greatest mistake made by
novice coaches is to translate the techniques and expectations of elite athletes too literally.
Young swimmers are less likely to need the inspirational motivation speech; the
high-powered technical analysis, or the complex training program that we (as coaches)
apply to our senior elite swimmers. Techniques for coaching juniors must be kept simple,
but within the realm of technical accuracy and professionalism.

Coaches of junior swimmers must also have a clear vision of what the end product of
their labour should be, but don't attempt to create a world record holder or Olympic
champion in the short-term. Try to create the best possible environment so that future
champions are allowed to develop. Very high training loads may create a fast swimming
10 or 11 year-old; but a well balanced program of technical skill development, physical
capacity improvement, and positive attitudes will provide the underpinning for the
highest levels of sporting achievement later in life. Teaching young swimmers good work
habits and responsibility for their training outcomes is an important part of this process.
Performing turns in accordance with swimming rules, streamlining off the wall,
controlling stroke and breathing pattern during all swims, maintaining predetermined
pace and technical form; these are all skills that must be taught and reinforced daily. The
overall environment in which training is conducted should be positive, competitive (yes,
children like to compete with each other during training) and full of feedback from the
coach.

The coach of junior swimmers should be mindful of how growth and maturation impact
upon performance. Often the performance improvements attributed to a training program
are primarily the result of growth and physical maturity. Training programs for junior
swimmers should be simple in their approach to physiological development. Young
swimmers need to progressively improve their aerobic capacity while also improving
upon natural speed. A balanced training approach will take this into account. Physical
preparation for current competition must be balanced with objectives for future
competitions. This means developing capacities that will enhance performance over a
range of racing distances during the formative years.
Chapter 1 Improving Coaching Skills Coaching Swimming: An
Introductory Manual (second edition)
7

The specific physiological demands upon children and mature-age swimmers are quite
different, even when the racing distance is the same.

Young swimmers need continuous mental reinforcement. This is accomplished in part by


maintaining an environment containing numerous opportunities for interaction with the
coach. In addition to verbal feedback from the coach, non-verbal communication is very
important. The coach must be able to provide reinforcement by his/her physical presence
(i.e. moving around the pool deck during training is desirable), using eye contact, and
demonstrating recognition of how each swimmer is performing. Although physical
training requirements may be simple for juniors, training requirements from the mental
perspective are more complex. Each day the coach must be able to express the core
training objectives in a slightly different and ‘fresh’ way so swimmers do not loose
interest. This may require considerable planning and lateral thinking on the part of the
coach.

Developing a healthy attitude toward competition is also a major objective for swimmers
training in junior squads. As mentioned above, it's natural for young swimmers to want to
compare themselves with each other. Healthy attitudes about competition include being
able to find a positive situation from every race experience, independent of a race
outcome (i.e. who finished first). In other words, the coach and swimmer must embrace
the process as much as the product. If a swimmer can improve his/her execution of race
skills, pacing and strategy, or any number of other performance components, then he/she
has ‘won’ in terms of personal achievement. The young swimmer must first learn to
judge him/her self in terms of individual potential. The inevitable comparisons with other
swimmers should be a secondary consideration, because each swimmer has little or no
control over the performance of other swimmers in a race. In this regard the coach is the
most significant role model for swimmers as they develop attitudes and behaviours
toward competition. The coach must also be sensitive to the influence that parents have
upon the attitudes of their children. Therefore, the coach assumes a role of counsellor or
educator to the parents as well as the swimmers.

Developmental Considerations
There are numerous accounts of how young swimmers progress from novice to elite. For
some individuals, performance improves at a steady rate and for others it advances and
declines in spurts. It's important for the coach to understand how some aspects of sporting
development are sensitive to change during certain periods of physical growth and
maturation. A physiological capacity such as aerobic endurance will develop quite rapidly
during the critical period of pre-pubertal growth. Another characteristic, such as muscle
power to body weight ratio, may decline during a period of rapid growth and limit
swimming performance temporarily. The coach should also be aware that certain skills
and physiological capacities interact. For example, stroke technique development will
affect the ability to swim efficiently and this will influence the acquisition of both
endurance and speed. It's evident that every child will mature at a slightly different rate.
Normative data give us an indication of what age to expect certain biological events to
take place. In reality, any two 10 year-old swimmers may be 1-2 years apart in their
biological ages. Physical training itself may impact upon the attainment of maturational
events. These points are made to

Chapter 1 Improving Coaching Skills Coaching Swimming: An


Introductory Manual (second edition)

8 emphasise the fact that variation between individuals must always be recognised by the
coach when planning a training program.

Age-Group Training Model

Four broad periods of age-group swimmer development have been identified (refer to
chapter 10, ‘Long Term Swimmer Development). A number of training objectives are
associated with each period based upon biological maturity, progressive skill
development, and emotional maturity. An age-group swimming program should deliver
outcomes in each of the target areas of development (i.e. training, motor learning,
knowledge and attitudes, competition) before the swimmer progresses to the next period.
To achieve all, or nearly all, of the objectives identified for each training period may take
several years of progressive training. Therefore, the coach should have a plan or vision of
how training will progress over time and from one period to the next. Squad composition
and training complexity will reflect what is appropriate for that level of maturation.
Individual variations to the model must be recognised. For example, a late maturing 12
year-old girl might best fit into the training plans developed for a group of mostly 10-11
year-olds. Conversely, a 12 year-old girl who has completed her growth spurt and passed
the age of menarche may be capable of meeting the training expectations of a more
advanced training period.

Three important concepts should be incorporated into the coach's planning. First,
different training periods should have different emphasis in terms of program delivery.
That is, the way you conduct the training program will vary according to maturational
considerations. Second, it's of paramount importance that both mastery and retention of
lower level skills and fitness components are carried forward. Swimmers must be able to
quickly re-acquire their prior level of fitness/skill after a period of reduced training.
Third, training plans should be progressive in their application, advancing in a logical
manner. Training plans build from year-to-year, some components remain the same and
increase in volume and intensity; other components are added when swimmers are ready
to absorb the effects of the training.

Summary of Key Points

SKILL DEVELOPMENT
• Good Technique = High Efficiency.
• It's important to know the difference between Technique and Style.
• Complex skills should be broken down into simplified technique.
• Skills develop from 'Learning' to 'Performance with Speed' to 'Performance with
Speed while Under Pressure'.
• The current level of physical preparation impacts on skill acquisition.
• Motor development and the complexity of a skill will affect learning.

PSYCHOLOGICAL DEVELOPMENT ♦ Maintaining motivation requires developing


psychological skills within the
context of the swimmer's rationale for participation. ♦ Age-group swimmers are not
‘little senior swimmers’ and senior swimmers are
not ‘bigger age-group swimmers’.

Chapter 1 Improving Coaching Skills Coaching Swimming: An


Introductory Manual (second edition)
9

♦ Success encourages high self-esteem. ♦ Positive reinforcement instils a sense of


success. ♦ The transition from age-group to senior swimming is psychological as well as
physiological. ♦ Psychological skills should improve with training and practice if applied
in a
positive environment. ♦ The emotional state of children is unpredictable and
can change quickly.

PHYSICAL DEVELOPMENT ∇ Success prior to the age of 12 is closely related to


biological development /
success at the senior elite level is the result of many factors. ∇ Late or average-age
maturing children tend to stay in the sport longer. ∇ Sporting preparation that coincides
with developmental stages will result in better
long-term improvements. ∇ Prior to the age of 12 the energy system that is developing
most rapidly is the
‘Aerobic System’. ∇ Appropriate endurance based training may be the single most
important
component of success throughout the career of an athlete. ∇ Energy systems must be
developed concurrently using progressive loading and
recovery methods. ∇ Positive experience motivates one to acquire higher
physical capabilities.
Chapter 1 Improving Coaching Skills Coaching Swimming: An
Introductory Manual (second edition)
10

CHAPTER 2 MECHANICS OF
SWIMMING

Biomechanics is the scientific discipline that describes and explains the forces acting
upon, or generated by, our body movements. Understanding basic scientific principles
helps the coach to analyse the effect of body position and movement on propulsive and
resistive forces. Every coach observes the way swimmers perform; by linking observation
with understanding (i.e. biomechanics) the coach can determine which factors contribute
to high quality performance and which may limit even better performance. Every coach
will develop a stroke model in each of the four competitive strokes that serves to
conceptualise an effective and efficient means of propulsion. Within that model there is
an acceptable range of variation of movement. Every swimmer develops a unique ‘style’
within the overall model for stroke ‘technique’. Variations occur because of individual
strength, flexibility, and anatomical considerations. Understanding the principles of
biomechanics will help the coach fine-tune swimming technique to suit the swimmer.

General Terminology

Swimming strokes have been described using many sets of terminology. For simplicity
this text will adopt commonly accepted terms that help to describe the movements
required to perform acceptable stroke patterns. All movements of the arms and legs are
divided into two phases: (1) propulsive and (2) recovery. Generally, we are concerned
with the positioning of the limbs to apply muscular force to generate forward propulsion,
but we must also prepare for the next stroke; so recovery movements are important to
overall stroke efficiency. Recovery movements are patterned so they keep resistance
factors to a minimum and allow the swimmer to maintain stroke rhythm and symmetry.
Movements of the head and trunk are also important because they allow the swimmer to
position the mouth so that rhythmic breathing can occur. Efficient trunk and head
movements allow water to flow smoothly over/around the body to aid streamlining.
Trunk position is critical to the effective application and transmission of force by the
limbs because all mechanical actions are part of a kinetic chain. The body's
musculoskeletal system of levers is used to transfer equal/opposite force from the water
through each body segment. This system of levers and subsequent transfer of forces is
known as the kinetic chain. Force is most effectively applied if there is a stable base of
support from which the mechanical actions are liked.

Propulsive movements are usually called sweeps because this terminology reflects the
combined applications of drag and lift principles of propulsion. The term is further
defined by indicating a direction of movement relative to the body, such as; outsweep,
insweep, upsweep, or downsweep. Sweeping movements are seldom uni-directional, so a
combination of descriptors such as insweep-upsweep is often used. Positioning of the
primary propulsive surface (i.e. hands and feet) is further defined by the angle or pitch of
the surface relative to the axis of the limb and the direction of movement. The hands and
feet are sensitive to pressure applied against the water, this allow the swimmer to position
the limbs precisely. The forearm and upper arm, as well as the segments of the leg, will
serve as secondary propulsive surfaces. The aim of efficient
Chapter 2 Mechanics of Swimming Coaching Swimming: An
Introductory Manual (second edition)
11

swimming is to present as large a surface area as possible to create a flow of water (i.e.
push-pull or speed of water flow) in a direction that yields a resultant force to produce
forward propulsion.

Some movements of the head and trunk are not propulsive, but they contribute to
streamlining by positioning the trunk for optimum transfer of force through the kinetic
chain. Trunk movements are described in relation to the body’s axes. In freestyle and
backstroke we allow the trunk to rotate around the long axis of the body. In butterfly and
breaststroke the axis of the trunk will rise and fall through the horizontal to create a
‘wave’ like motion. All body movements may be analysed with respect to five basic
elements of technique: (1) streamlining or body position, (2) kicking, (3) armstroke, (4)
breathing, and (5) timing and co-ordination.

The coach should develop a model for each swimming stroke that becomes a frame of
reference for teaching and refining technique. Deviations from the model should prompt
the coach to identify if problems exist and then propose strategies for correction.

Buoyancy and Gravity

These are opposing forces which act passively through a point in the body. If the
opposing forces act through points which are close together, then the body will float in a
roughly horizontal position. If the opposing forces act through points which are apart; a
rotation of the body’s axis will result. The force exerted by gravity will push the body
downward, but this is counteracted by the upward force equal to the displacement of
water; called buoyancy. The relative magnitude of these opposing forces will change
during the stroke based upon the distribution of body mass. Every time a body part is
lifted above the surface of the water, the gravitational force becomes greater. This is
particularly important when the head is lifted to take a breath in breaststroke and butterfly
or the arms are lifted to facilitate the recovery phase of a stroke. There are both
advantages and disadvantages related to the overall resistance force created when a body
part is lifted out of the water. Holding bodyweight out of the water will increase the
downward force; as an example, a butterflyer who keeps the head up will find it hard to
keep the body in a streamline (i.e. horizontal) position because gravitational force acting
through the body will push the legs down. However, lifting the arms out of the water is
the only way to facilitate recovery in freestyle, backstroke, butterfly, and is also a
common recovery method in breaststroke. In this case lifting a body segment out of water
will decrease the water resistance created by limb movement. Because the limbs are
moving in a direction that can not produce propulsion during the recovery phase, an out
of water recovery results in greater efficiency.

As a swimmer grows and matures the size and composition of body tissues change.
Increased bone and muscle mass improve the potential to generate propulsive force, but
make the swimmer more dense or heavy in the water. At the same time increased lung
volume and body fat make the swimmer less dense or lighter in the water, but may
decrease (particularly if body fat increases are substantial) the body’s power to weight
ratio. Swimming technique is influenced by the relative balance of these

Chapter 2 Mechanics of Swimming Coaching Swimming: An


Introductory Manual (second edition)

12 simple forces as the body matures and changes. Any single change results in numerous
trade-offs in other factors that influence propulsive force and resistance.

Inertia and Momentum

Inertia is the amount of force that must be generated to set a body (or body part) into
motion. When we look at whole body motion, it’s clear that the summation of forces from
the limbs must be directed so that forward propulsion results. To overcome inertia the
propulsive force must be greater than the total of all resistance forces. Once inertia is
overcome, momentum is the product of body mass and velocity. Swimming efficiency
increases if we can conserve as much forward momentum as possible, this is done in
several ways. A streamlined body position helps to reduce the amount of water resistance
and thus, maintain forward momentum. By maintaining a stroke rhythm that balances
right-left arm/leg movements a swimmer voids stop-start actions and conserves more of
the forward momentum generated. In butterfly, the double arm recovery motion becomes
a continuation of the propulsive out-upsweep at the finish of the underwater stroke. Force
generated during the final stages is carried through to the recovery; stopping the arms as
they exit the water would break forward momentum. In breaststroke, the feet recover by
moving up toward the buttocks while the thighs remain angled back; this helps to
streamline water flow under the body to conserve forward momentum while the arms
recover forward. Every stroke has examples of how propulsive movements are translated
into recovery movements and streamlining techniques are used to conserve forward
momentum.

Sometimes momentum is not generated in a linear direction. The example of the butterfly
recovery is a good one. The wide sweeping action of the arms will develop momentum
(i.e. the product of the mass of each arm and its velocity) that is translated in an angular
direction. However, because both arms move simultaneously the angular momentum from
each arm cancels out the other. Wide sweeping single-arm recovery in freestyle or
backstroke is not used because it would generate a lateral force that may push the legs
from side to side. The correct way to use the angular momentum generated during a
backstroke or freestyle recovery is to align the direction of movement with the long axis
of the body. Rotation of the trunk positions the each arm so that it travels in an arc above
the body’s long axis. Angular momentum is also used during tumbling or pivoting
movements such as turning. Efficient turns are executed in a ‘tuck’ position because it
concentrates a greater amount of mass closer to the centre of rotation and this increases
the speed of rotation.

Levers and the Kinetic Chain

The joints of the shoulders and hips must apply leverage against the trunk to produce
forward body propulsion. However, for this to occur the force applied by the hands and
feet must be transferred through intermediate joints (i.e. wrist and elbow, or ankle and
knee). The joints of the limbs form a chain to link the transfer of force. As mentioned
before, the propulsive force must be sufficient to overcome the inertia of the body and all
external resistance factors. The body’s mass is levered forward across joints that act as
the fulcrum of the lever system (i.e. joints are the pivot point in the lever set-up). Because
the skeletal structure of humans places the fulcrum very

Chapter 2 Mechanics of Swimming Coaching Swimming: An


Introductory Manual (second edition)
13

close to the weight being moved, the resistance arm of the lever (which is the propulsive
surfaces of the arms or legs) must move through a relatively long distance. In all
swimming strokes it’s important to position the resistance arm so that force can be
applied along its full length. This means that at the beginning of each stroke as the
shoulder is still extended the resistance arm is defined by the distance from elbow to
fingertips. The classic ‘high elbow’ catch position allows the full length of the forearm
and hand to apply force. In freestyle, backstroke, and butterfly the resistance arm of the
lever system will increase in length as the hand(s) move deeper. As the hand(s) move
under the body (in the case of backstroke the hand moves to the side of the body, but
trunk rotation also brings the hand closer) at mid-stroke the lever system can apply
maximum muscular force. The arm(s) extend back, pushing water until the transition
from propulsive phase to recovery phase is complete. The hand must then move upward
toward the surface, this reduces the length of the lever’s resistance arm to the length of
the hand. Throughout the propulsive phase, force generated by the arms and legs is
transferred through the kinetic chain to the muscle attachments on the trunk. The trunk
becomes the stable platform and the body is levered forward.

As the length of the resistance arm gets longer there is a larger surface available to push
water backward, and thus creating forward propulsion. However, the musculature of the
shoulder and trunk must be equal to the load created. Swimmers who are not as strong
may compensate by simply shortening the length of the resistance arm. This is why some
swimmers use more elbow flexion, combined with inward rotation of the upper arm, to
position the hand(s) under middle of their trunk in freestyle and butterfly. Other
swimmers maintain a straighter arm position or have less internal rotation to position the
path of their armstroke under the shoulder. Breaststroke uses the same basic leverage
principle, but because the underwater stroke is much shorter there are actually two
distinct propulsive movements. The outsweep in breaststroke increases the length of the
resistance arm and positions the propulsive surface to push water back. The insweep is a
powerful movement that develops hand speed which produces more of a ‘lift’ or sculling
effect. As with the other strokes, the propulsive phase in breaststroke will have a smooth
transition into the recovery phase.

Velocity and Acceleration

Velocity is the directional speed of movement. Increased velocity of the limbs will
produce greater propulsion only if the surface is positioned to maximise the application of
force. Swimmers sometimes make the mistake of trying to move the arms and legs ‘fast’
without regard to the position and application of force. One of the most common stroke
faults is the ‘dropped elbow’ during the first half of a stroke pattern (note: the dropped
elbow syndrome can occur in any stroke). The swimmer has the sensation of moving the
hand fast, but the primary and secondary propulsive surfaces (i.e. hand and forearm) are
not in a position to push water back. Water slides off the surface of the arm and effective
propulsive force is reduced. Increased limb velocity is always achieved at a cost, the
amount of energy required to produce faster muscle contractions goes up dramatically. A
number of factors must be considered when coaching decisions are made with regard to
limb velocity. Sometimes a slower stroke is required to maintain optimum contact with
the water, at other times a faster stroke is advantageous. The coach must always
remember that whatever stoke factors

Chapter 2 Mechanics of Swimming Coaching Swimming: An


Introductory Manual (second edition)

14 are desired under race conditions must be trained; the body must adapt to the relevant
technical and physiological conditions.

Acceleration is a positive change in velocity and is an important factor in every


swimming stroke. The speed of arm/leg movement continually changes during the
propulsive phase of a stroke cycle. This happens because several mechanical principles
interact. First, force is required to overcome inertia and set the limb in motion. Second, at
various positions in the stroke cycle the mechanical application of force is either limited
or optimal due to the positioning of the lever system of the limbs. Third, because force
application is linked through the kinetic chain, certain body positions and movements will
contribute to limb acceleration. In general, the ability to apply muscular force through the
limbs is less when the propulsive surface (i.e. hand) is at its' most distant point from the
trunk; this is at the start of each stroke. During each stroke mechanical factors change as a
consequence of resistance forces, limb and body position, and streamlining factors.
Theoretically, it would be energy efficient to apply force evenly throughout each stroke
cycle; however, this is impossible and does not reflect the complexity of human aquatic
movement. The hand generally accelerates to a point of maximum propelling force, this is
usually reached about half to two-thirds of the way through the freestyle, backstroke, or
butterfly propulsive phase. In the breaststroke there are two peaks, one near the end of the
outward hand movement and one near the end of the inward hand movement.

Continuity of Motion

The development of peak propelling force (i.e. through hand acceleration and optimal
positioning of the kinetic chain) must be balanced with sustained propelling force. From
observations on elite swimmers it appears they 'hold the water through the stroke' much
better than average swimmers. Studies on elite and sub-elite swimmers of similar size and
strength actually show very little difference in the amount of peak force produced by the
hand. But great differences exist in the total force production during the stroke cycle.
Elite swimmers apply more effective force throughout the stroke because they have
greater stroke length and fewer low points of force production during each stroke cycle.
Certainly muscle strength, hand speed, and limb size are important, but these
characteristics must be combined with the optimum application of force through correct
stroke technique. Correct technique (other factors being equal) is the single most
important characteristic of successful swimmers because it increases the level of
swimming efficiency.

Another principle that contributes to efficient swimming is the energy cost of movements.
We try to maximise movements that contribute to forward propulsion and minimise
movements that increase resistance factors. Efficient stroking is smooth and co-ordinates
the concurrent or alternating movement of body parts. Stroke faults develop when
stop-start or poorly synchronised or timed movements are used.

Propulsive Forces

Propulsive forces in swimming must be thought of in a number of ways. Movement of the


arms and legs results in a corresponding movement of water in the opposite

Chapter 2 Mechanics of Swimming Coaching Swimming: An


Introductory Manual (second edition)
15

direction; this produces a simple action-reaction force. However, unlike the forces applied
during land-based movement, the action-reaction in water is not against a fixed surface.
Movement of the limbs against the water causes the water to move, in the process only a
portion of the reactive force generated actually contributes to forward propulsion.
Although pushing against the water is our primary source of generating propulsive force
(this is called ‘drag propulsion’), it is not the only way of producing enough force to
move the body forward. The speed and direction of water flowing across the propelling
surfaces of the body will change the pressure created against those surfaces. Water
striking the leading edge of a moving surface is deflected around the surface. The
direction and angle of the surface can create a difference in the speed of water flow which
creates a pressure difference. As with our application of push-pull force, the transfer of
force through the kinetic chain allows the trunk to move forward. The lift theory explains
why swimmers can move their hands and feet in vertical or lateral directions and propel
themselves forward. The best example of this application of force is the technique of
sculling used by synchronised swimmers. With the body in a prone or supine position,
sculling movements similar to a figure '8' pattern will move the body. Conceptually we
envision arms and legs pushing against still water (i.e. not moving). However, water-
flow over the propelling surfaces, along and around the body, creates a dynamic
environment where fluid pressure can create lift force. In recent years there has been
vigorous debate over whether one type of propulsion is dominant. Researchers currently
suggest that drag propulsion is the major contributor, but both sources of propulsion
should be important because the contribution of each to the total propulsive force is
constantly changing.

Resistance Forces

Resistance forces counteract propulsion in several ways. First, because water is a dense
fluid (i.e. relative to air) it will resist displacement. A swimmer must push water out of
the way to move through it; this is known as frontal resistance. Second, the motion of
water next to the body creates a frictional force, this is known as surface drag. Third,
water being pushed away from the body or flowing off/around the body will collide with
still-water and become turbulent; this is known as eddy resistance. Obviously, the shape
of a body moving through the water becomes an important factor in terms of the amount
and direction of water displacement and ultimately the amount of resistance to swimming
movement. When large amounts of water must be displaced, or when the deflection of a
mass of water is very abrupt, the resistance is high. In general, it's more efficient for
water to flow smoothly in straight or gently curved pathways than to become turbulent by
changing direction sharply. Frictional resistance against the body is minimised by tight
fitting swimwear, swim cap, or shaving body hair. Turbulent water flow is minimised by
maintaining streamlined shape and body position so that water is not deflected sharply.
Efficient timing and coordination of stroke components will reduce water resistance
substantially, as well as reducing the energy cost of swimming at any particular velocity.

Chapter 2 Mechanics of Swimming Coaching Swimming: An


Introductory Manual (second edition)

16 Explaining Swimming Movements

To explain the push-pull force a swimmer applies, we turn to Newton's Third Law of
Motion, which tells us that for every action there is an equal and opposite reaction. In
other words, pushing water in one direction causes a reactive force in an opposite
direction. Several factors become important considerations when evaluating the
swimming strokes: (1) the size of the surface area pushing water, (2) the direction and
distance water is moved, and (3) the speed of the pushing surface. Together these factors
determine the relative contribution of each type of force to the overall propulsion.
Swimming is a relatively inefficient form of movement because a large proportion of
muscle force is wasted in non-propulsive applications. We commonly say that water
‘slips’ off the pulling surface. Some of the muscular force is also lost in the transfer
through the body's segments. In addition, some applied force is necessary to correctly
position the body rather than contribute to forward propulsion.

The study of stroke mechanics has highlighted the need to maximise propulsive forces
while minimising resistance forces. One way of applying force over a longer period of
time within a stroke cycle is to adjust the depth, width and length of the stroke pattern.
This allows the hand to find 'still water' and impart a reactive force on it. This application
of technique also allows the hand to build up speed during the underwater phase of the
stroke cycle and thereby increase the potential that lift force will contribute to propulsion.
The hand(s) must move the largest possible volume of water. However, as with any
movement there is an optimal pattern. Force application is more difficult to maintain
when the hand moves too far from the body. Also, if joint angles do not transfer force
effectively (i.e. the classic elbow drop), the reactive force is angled downward instead of
in a direction parallel to the long axis of the body.

Forces are also transferred through the body during the recovery cycle. When the arms
are recovered over the water (i.e. freestyle, backstroke, butterfly) the distance travelled,
weight distribution from the point of rotation, and limb speed are all factors that affect the
way force is transferred through the body. For example, knowledge of how force is
applied and transferred during the recovery phase of freestyle suggests that by shortening
the distance between the centre of mass of the arm and the point of rotation (i.e. shoulder
joint) we will reduce the amount of force necessary to mover the arm forward. Therefore,
a bent elbow recovery is an effective technique, but must be coordinated with the correct
trunk rotation for optimum efficiency. We often see highly skilled swimmers using an
alternate technique that may not seem efficient. One example is the straight arm or
‘windmill’ like recovery used by some freestylers. Some swimmers are able to use this
technique effectively because their body rotation, shoulder flexibility, timing and
muscular control allow them to keep lateral forces to a minimum. This is another example
of a style variation within the stroke model.

During the backstroke arm recovery there is also a transfer of momentum. In this case it's
an advantage to keep the arm straight during the course of the recovery movement. As the
arm is lifted from the water some muscular effort is required; however, once the arm
reaches the top of its arc the weight of the arm itself carries it downward in a ballistic
movement. Backstrokers should try to keep their arm recovery in line with the long axis
of the body, avoiding a recovery pattern which swings the arms to the

Chapter 2 Mechanics of Swimming Coaching Swimming: An


Introductory Manual (second edition)
17

side. As with freestyle, body roll assists the desired movement by allowing forces to act
parallel to the long axis of the body.

We tend to visualise movement of the arms and legs relative to the swimmer's body. This
is actually a good technique used by coaches to explain the strokes and give a visual
representation of the movement patterns. This is called relative movement, because it's in
perspective to the body. However, we know that the body is constantly moving forward.
When the hand movement is viewed in relation to its' actual directional path (i.e. in
relation to a stationary object, such as a point on the side of the pool), the shape of the
pulling pattern may appear compressed or distorted. Coaches studying illustrations or
photos of stroke mechanics should clearly understand whether relative or actual
movement is being depicted and from what viewing perspective (i.e. bottom, side, front).
Chapter 2 Mechanics of Swimming Coaching Swimming: An
Introductory Manual (second edition)
18

CHAPTER 3 FREESTYLE and


BACKSTROKE
Freestyle – Overview

Technically speaking, 'freestyle' swimming can take the form of any style. However, the
fastest method currently devised is the Australian Crawl and therefore, freestyle has
become synonymous with crawlstroke swimming. Freestyle is the fastest of the four
competitive swimming strokes because the alternating arm action and continuous kick
provide continuous application of propulsive force. Resistance forces are minimised
because the body's long axis remains parallel to the direction of movement. These two
conditions exist during backstroke swimming as well, but the prone body position used in
freestyle allows the musculature of the chest and upper back to pull the arms through the
water with a greater application of power through the kinetic chain. A prone body
position also allows the hand-arm surface to be positioned immediately after hand entry
so that propulsive power is generated through a long stroke cycle. Backstrokers take a
fraction of a second to position the hand-arm effectively at the start of each stroke cycle.

Streamlining the body during freestyle swimming is achieved by maintaining (as best as
possible) a long, straight and slender shape. Movement of water around the body tends to
be in parallel streams or layers and this flow will be broken if excessive lateral motion of
the head, shoulders, or hips occurs. There will be some turbulence created by the kicking
action, but because this turbulence is at the trailing end of the flow it presents less of a
problem. The propulsive force generated by the kick serves a very important function – it
helps to maintain trunk stability and body position. Streamlining of the trunk is also
assisted when the shoulders rotate as each arm extends forward for the hand entry. A
smooth rolling action of the trunk, approximately 40-45 degrees to either side of the
long-axis of the body also assists in positioning the arm for a smooth recovery. Rolling of
the trunk combined with turning the head to the side is used to facilitate breathing without
altering the body's position from the horizontal. There is some resistance created from the
frontal contact of water at the head and shoulders, but this can be kept to a minimum if
the head is comfortably aligned with the body. In fact, the small wave produced around
the head is used to the swimmer's advantage when breathing into the trough that follows
the bow- wave.

Optimum efficiency in freestyle swimming relies upon the application of propulsive force
that is both continuous (i.e. no gaps between impulses) and prolonged (i.e. propulsive
force is applied through the full range of motion). This is combined with good
streamlining techniques that minimise active drag.
Chapter 3 Freestyle and Backstroke Coaching Swimming: An
Introductory Manual (second edition)
19

Kick

1. Heel of the left foot lifts to the surface, then 2. As the left leg moves down the right leg
knee bends and leg pushes down. stretches up.

Kicking action consists of alternating propulsive and recovery movements of the legs,
known as a flutter kick. Beginning swimmers should focus on establishing a strong kick
because this helps to stabilize the body and maintain streamlining. As the swimmer gains
in strength and skill the armstroke will assume a greater proportion of the overall
propulsive force generated. However, the role of kicking remains an important
component to efficient and effective (i.e. fast) freestyle swimming.

The kick begins with hip extension to lift the sole of the foot to the surface. The knee
remains flexed as the leg begins to drive downward. Knee extension completes the
movement in a whip like action; water is pushed backward and downward and slightly
inward off an extended foot position. The ankle remains in an extended position
throughout the kicking sequence to lengthen the kicking surface. The simultaneous
recovery action of the opposite leg involves extension at the hip and knee. As the upward
movement takes the leg past the horizontal axis of the body the knee begins to flex
slightly in preparation for the next downward propulsive movement. Typically 2 or 6
beats (i.e. each downward movement being one beat) are performed during each stroke
cycle. Six beat kicking appears to be continuous, there is no break in the alternating
movement of the legs. A 2-beat kicking tempo has a slight pause in the leg movement
when right-left feet are at the opposite extremes in the range of motion. A fast kicking
tempo drives the hips upward and keeps the trunk positioned high in the water, but this is
achieved at a greater cost in terms of energy consumption than a slower kicking tempo.
Energy consumption is the reason most long-distance swimmers prefer a slower kicking
tempo. Regardless of the kicking tempo, the finish of an armstroke will coincide with the
downbeat of the leg on that side of the body to facilitate the roll of the hips and a smooth
transition to the arm recovery.

A characteristic feature of the two-beat kick is the “split” position of the legs at the end of
each downbeat; this serves to stabilize and streamline body position. It’s often the case
that two-beat kickers perform smaller kicks between their two major ones; this gives the
impression of a four-beat kick. However, these swimmers are usually considered to be
two-beat kickers because the additional beats do not produce much effective force. The
mechanics of the two basic kicking tempos, 6-beat and 2-beat are identical, only the
tempo differs. It’s worth noting the tendency for distance swimmers to use the slower
tempo during some stages of a race and also have the

Chapter 3 Freestyle and Backstroke Coaching Swimming: An


Introductory Manual (second edition)

20 ability to change to a 6-beat tempo as required. Virtually all sprint swimmers use a
strong six-beat kick.

Armstroke – Propulsive Phase

The pulling pattern, is often described as being shaped like a narrow and elongated ‘S’
when viewed from under the body. This is a generalization, because individual pulling
patterns of skilled swimmers may vary from being almost straight (i.e. very little lateral
directional change) to a distinct outsweep-insweep during the underwater cycle. As the
hand enters the water the trunk should be rotating to that side, the elbow remains higher
than the wrist as the forward momentum of the arm allows the hand to extend forward
smoothly and the fingertips begin to point downward. Some swimmers reach full (or
nearly full) arm extension with the hand just under the surface; while other swimmers
move the hand downward more sharply to begin feeling pressure on the water from a
slightly deeper position. When pressure is felt on the hand-forearm the swimmer has
begun the 'catch' phase of the stroke. From the catch position the hand sweeps down (or
down and slightly outward) while the elbow is maintained in a high stable position. The
elbow position is critical at this point because any drop or backward slip will reduce the
swimmer’s ability to produce optimum propulsive force. The first photo of the male
swimmer (see above in the discussion of kicking) shows this high elbow position early in
the stroke when the hand is forward of the shoulder.
The hand may come under the trunk during the middle of the pulling pattern. Hand
position may vary from directly under the shoulder to across the trunk and under the
opposite shoulder; anything within these limits fits within the freestyle stroke model. Not
every freestyle swimmer will have the same amount of bend at the elbow; a ‘normal’
range of angles might be nearly straight to about 90-degrees at mid-stroke. The female
swimmer pictured below demonstrates a bent arm position whole the male demonstrates a
straighter arm position.∗ Swimmers who tend to be relatively straight with the arm will
naturally have a deeper pulling pattern and usually have less lateral movement of the
hand. This usually means the hand position at mid-stroke is under the edge of the trunk.
These differences represent acceptable variation in the stroke pattern as long as the
swimmer has the strength to hold the elbow in line with the wrist (as viewed from the
side) and not let the elbow position slip backward. Some coaches associate a straighter
arm position with sprinting; however, this may not be the case. Many middle-distance and
distance swimmers maintain an elbow position much greater than 90-degrees. Stroke
technique is not gender specific and variation may occur as each swimmer seeks to
maintain propulsive force within his/her own limitations of strength and flexibility.

From the deepest point in its' range of movement the hand sweeps upward, or slightly
outward-upward, as it passes the hip. The amount of both inward and outward movement
of the hand will vary between any two swimmers and may also be slightly different from
right to left hand in any one swimmer. Naturally, approximate stroke symmetry (i.e. right
and left arms as mirror images) is desirable because it indicates a balance in the
movement pattern on each side of the body.

∗ Some swimmers may bend their elbow even less than the male swimmer pictured.

Chapter 3 Freestyle and Backstroke Coaching Swimming: An


Introductory Manual (second edition)
21

Hand sweeps under the trunk, elbow bend about Hand stays in line with the shoulder, elbow 90
degrees. bend is much less.
Hand velocity should increase (i.e. acceleration) through the second half of the pulling
pattern. Muscle strength and power, as well as arm position, determine the amount of
hand acceleration achieved. Hand acceleration, provided the pulling surface of the hand
and forearm are kept in a position to hold the water, is a key factor in attaining peak
propulsive force. The rolling action of the trunk also adds power to the final phase of the
stroke.

Armstroke – Recovery Phase

The recovery phase of the armstroke actually begins slightly before the hand leaves the
water. When the arm has reached almost full extension the palm surface of the hand will
naturally turn inward. A lifting of the elbow allows the hand to slide out of the water with
minimum surface resistance. Rotation of the body around the long- axis facilitates arm
recovery with a minimum of lateral force transferred through the trunk. A high elbow
position during the mid-stage of recovery keeps the forearm in a relaxed position with
fingertips pointing down. The forward extension of the arm is a relaxed swing; flexibility
and muscle control around the shoulder girdle are required to make the movement smooth
and continuous. The degree to which the arm is bent or straight during the recovery will
have an affect on the transfer of force through the body because it influences the distance
of the hand from the trunk and thus the amount of rotational force that develops. Each
technique variation has advantages and disadvantages. Shoulder flexibility also becomes
a consideration in the recovery pattern adopted by each swimmer. A higher degree of
muscular control is required if the straight arm recovery (i.e. windmill action) technique
is used. A bent-elbow technique allows for a smaller radius of rotation and less muscular
effort is required to swing the arm forward.

Breathing
Breathing is accomplished by turning the head to one side with the natural roll of the
trunk. It's important that the turning motion is accomplished with very little lifting of the
head because this may disrupt body position or interfere with the rhythmic motion of the
armstroke. One ear remains in the water when the mouth breaks the surface. Inhalation
begins as soon as the mouth breaks the surface because expulsion of air has

Chapter 3 Freestyle and Backstroke Coaching Swimming: An


Introductory Manual (second edition)
22

taken place when the face is submerged. A breath is normally taken during every
complete stroke cycle. However, most swimmers find that it's an advantage to limit
breathing somewhat during sprint events and in certain race situations; such as one or two
strokes in/out of turns and the first and last few meters of a race. This seems to help lift
the stroke-rate temporarily and provide an advantage when racing. In events longer than
50m swimmers should not be encouraged to hold their breath for long periods. Normally
a breath is taken every 1-3 seconds while swimming. Depending upon the stroke-rate this
means one breath every third stroke is sufficient to supply enough oxygen for sustained
swimming. Swimmers using an average stroke-rate will not have difficulty breathing
bilaterally (i.e. one breath per 1 1/2 stroke cycles). A few research studies have looked at
the kinematics of the body roll and found no significant differences in the propulsive
forces applied during breathing and non- breathing stroke cycles. Studies investigating
the changes in resistance forces during breathing and non-breathing strokes are more
varied. It appears that when a swimmer incorrectly times the breath (in relation to where
it should occur in the stroke cycle) there may be a sharp increase in resistance forces
acting on the body.

Timing and Rhythm

Because the arms do not move at a uniform speed within each stroke cycle, it's important
for every swimmer to develop a rhythm. Deviations from the stroke model that break the
natural rhythm of a stroke may include these factors: breathing late during the arm
recovery phase; incorrect timing of the downbeat of the kick; excessive overlap of
right/left arm propulsion; erratic changes in the length of each stroke cycle. All of these
technical variations will reduce swimming efficiency. Swimmers often change the timing
of stroke components when trying to swim at a higher or lower stroke frequency (i.e. very
fast or slow), this is a mistake. The co- ordination of stroke components (i.e. body roll,
kicking tempo, pulling pattern, and breathing) that a swimmer wishes to use during
competition should be practiced at all times and under all training speeds. The stroke-rate
used at any particular swimming speed will vary, but the rhythm of the stroke should not.

Backstroke – Overview

Backstroke rules allow any type of arm and leg movement; the only requirement is that
the body must remain on the 'back'. If the shoulders rotate past vertical, or 90- degrees to
the surface of the water, the swimmer is no longer on the back. The most efficient
backstroke is a backcrawl technique that uses a flutter kick (except during the start and
turns when a dolphin kick is often used) and an alternating armstroke.

Streamlining is similar to freestyle; the


ideal body position is as close to horizontal
as possible. This maintains a smooth flow
of water over/around the body. Resistance
is increased if the kicking action lifts the
knees out of the water or if the hips are
allowed to drop.

Chapter 3 Freestyle and Backstroke Coaching Swimming: An


Introductory Manual (second edition)
23

Head position will also have great affect on streamlining. If the head is held too high, or if
lateral movement is allowed, the trunk may deviate from the horizontal. These actions
result in a body position that increases frontal resistance. The head should be kept
comfortably aligned with the body in the horizontal plane, independent of body roll. A
smooth rolling action of the trunk, approximately 45 degrees to either side of horizontal,
will assist in rotating the shoulders for both arm recovery and application of propulsive
force.

Kick

The leg action is similar to freestyle. Because each arm recovery is made using a high,
out of water motion, there are considerable downward forces acting against the horizontal
positioning of the body. Therefore, a strong 6-beat kick must be used to keep the hips
near the surface, as compared to freestyle swimming where a 2-beat kick is often used.
The flutter kick used in backstroke swimming relies upon the upbeat as the propulsive
phase. Hip extension moves the leg downward (recovery phase) from the horizontal until
the foot reaches its lowest point. The knee bends to position the lower leg and foot and
then a combined hip flexion and knee extension drives the foot to the surface, ending in a
whip-like action. As the foot drives upward it also angles inward slightly, finishing with
the toes breaking the surface. Ankles are extended to increase the surface area of the leg.
Correct mechanics of the kick will assist the roll of the hips and trunk.

Armstroke – Propulsive Phase

The pulling pattern is best described as being shaped like an ''S' on its' side. The arm is
straight and extended behind the shoulder as the hand enters the water. The shoulder will
roll substantially toward the entry arm. The hand is turned so that the palm is facing
outward upon entry, this allows the hand to enter on its little finger edge to minimise
resistance and maintain the momentum developed during the arm recovery. The hand
moves downward and somewhat outward until it achieves the desired depth. At this stage
the fingers may be pointed either slightly downward or to the side. Pressure increases on
the hand as it begins to apply force. As with freestyle the catch position is defined as the
first point at which propulsion is generated. There is some debate amongst experts
regarding where the catch actually begins. Some believe that lift forces begin to act as the
hand moves downward, others argue that the catch begins when drag forces take effect as
the hand reaches its maximum depth. Each argument may be correct, as the depth
attained, distance the hand travels away from the body, and angle of the hand all affect
the magnitude and direction of the propulsive force generated.

From the catch position an upsweep of the hand brings it close to the surface at mid-
stroke. This action is combined with a slight insweep due to increasing elbow bend. An
angle of approximately 90-120 degrees between the forearm and upper-arm is

Chapter 3 Freestyle and Backstroke Coaching Swimming: An


Introductory Manual (second edition)
24

reached at mid-stroke. As with freestyle, the elbow must be held in a position, relative to
the hand, so that it does not slip. If the elbow leads the hand, maximum pressure will not
be maintained on the hand-forearm surface. From the point where maximum elbow bend
is achieved, the hand pushes back and downward as the arm extends. Hand speed
increases during this downsweep because the hips are rolling. During the final third of the
stroke pattern the wrist flexes so the hand continues to maintain pressure on the water.
When the arm is fully extended the palm turns inward and the hand begins to move
upward as the shoulder lifts to start the recovery phase. Slight right-left differences in
symmetry are common and do not necessarily represent a problem; flexibility
characteristics and muscle strength are not always perfectly balanced.

Armstroke – Recovery Phase

During the recovery the arm remains straight, elbow firmly fixed, as it swings upward in
an arc directly above the shoulder. Arm recovery is assisted by rotation of the trunk
around the long-axis of the body. When the arm reaches a vertical position and begins its
downward arc, the hand should rotate so that the palm is facing outward in preparation
for a smooth and streamlined entry with the edge of the little finger leading. The timing
of each underwater stroke and recovery keeps the hands at opposite sides of the stroke
cycle. Rolling also allows the arm entering the water to smoothly slide downward to
achieve sufficient depth for the catch. Breaks in timing may occur when the shoulders do
not roll sufficiently or the head moves laterally. Efficient hand entry is critical because it
helps to maintain the momentum generated by the arm during its recovery movement. If
the knuckle side of the hand strikes the water flat upon entry it slows this downward
movement of the hand to the catch position. The ideal arm recovery movement is through
a 180-degree arc directly over the shoulder. Any hand/arm movement across the long axis
of the body will result in a lateral transfer of force through the trunk. If the hand enters
the water past the mid- line of the body it’s very likely the elbow will bend and a smooth
transition from recovery to catch position will be broken.

Breathing

Because the swimmer is on the back, with the mouth out of water, it's easy to overlook
breathing as a consideration when evaluating technique. However, it's important for
backstroke swimmers to maintain an even breathing rhythm. A good technique is to
inhale every time the left (or right) arm recovers and exhale as the opposite arm recovers.
Inhaling and exhaling during backstroke may be a little more

Chapter 3 Freestyle and Backstroke Coaching Swimming: An


Introductory Manual (second edition)
25

explosive than during freestyle. Swimmers should try to take a full breath each stroke
cycle and avoid shallow rapid breathing. During the underwater segment following the
start and every turn, it's normal to exhale continuously from the nose and mouth to keep
water out of the nose.
Chapter 3 Freestyle and Backstroke Coaching Swimming: An
Introductory Manual (second edition)
26

CHAPTER 4 BREASTSTROKE and


BUTTERFLY

Breaststroke – Overview

Breaststroke is the least efficient of the four competition strokes because of the large
amount of water resistance created due to body position and an underwater arm recovery.
FINA rules stipulate that the elbows must remain underwater during the recovery
movement. Even though some swimmers recover the hands slightly above the surface, the
forward movement of the arms is primarily underwater, and thus resistance is high. The
width of the kick also influences body position and streamlining. However, development
of the ‘wave technique’ has allowed some breaststrokers to achieve improved
streamlining. When compared to the other strokes, breaststroke technique uses very
precise timing to balance the contribution of arm and leg propulsion as well as streamline
the upper and lower body segments. Although some swimmers may still rely on an
exceptional kick for peak propulsive force the breaststroke model should consider both
arms and legs in the propulsion- resistance equation.

Streamlining

There are three key phases of the stroke where streamlining of all, or part, of the body is
critical. First, at the beginning of each stroke cycle the body will be completely
streamlined with the head or face submerged. Arms are extended, hands together, and line
of sight is down-and-forward. Hips are at, or just below, the surface with the legs fully
extended and toes pointed. The second phase of streamlining occurs as the insweep phase
of the armstroke is completed. Because the insweep is a very powerful propulsive action,
the trunk is moving forward. At this instant the leg recovery is just beginning. If the
swimmer starts drawing the heels toward the buttocks too early, some of the forward
momentum will be lost. The shape of the body at this point in the stroke allows water to
travel smoothly backward without being deflected sharply at the thighs. Streamlining the
legs in this way helps to maximise the forward propulsion by reducing the negative
influence of resistance against the legs.
Chapter 4 Breaststroke and Butterfly Coaching Swimming: An
Introductory Manual (second edition)
The third phase of streamlining occurs as the legs drive backward; this is the propulsive
part of the kick. Streamlining of the upper body is achieved because the head is lowered
between the extending arms and the trunk is stretching forward. Because the trunk is
following the arms forward the hips will tend to lift as the feet come together at the finish
of the kick. Breaststrokers using a ‘wave’ style stroke will lunge forward with the upper
body high in the water (this is often combined with an arm recovery slightly above the
surface) to allow the hips to lift.

Armstroke – Propulsive Phase

The armstroke generates propulsion from both an outsweep-downsweep and insweep-


upsweep of the hands and forearms. From a full streamlined position the hands move
outward with the palms angled slightly. The initial application of propulsive force, as the
hands scull outward, probably results from greater application of lift forces than drag (i.e.
push) forces. Some breaststrokers begin the armstroke with the hands slightly deeper at
the catch position, their first movement of the hands may be slightly upward as well as
outward. As the hands move outward the elbows remain in a high position and begin to
flex so the hand-forearm position can apply greater drag propulsion. At the widest point
of the outsweep a breaststroker’s arm position resembles the pattern used during butterfly.
During the outsweep the hands and elbows remain forward of the shoulders, if the elbows
slip back efficiency will be reduced. Because the hands must reverse their direction very
sharply, some propulsive power is lost during the transition. Once the directional change
of the hands is made from out-to-insweep, the relative contribution of drag force will
increase. Drag force is perhaps more important to breaststroke propulsion than previously
thought. Forceful adduction of the shoulders directs the hands on their inward-upward
movement pattern. This also lifts the head and shoulders out of the water. As the hands
come together the elbows are positioned under the outer edges of the shoulders and are
pointing downward. The hands travel through a loop-like pattern when viewed from the
front. Arm recovery is a smooth transition following the insweep. Some swimmers start
moving the hands forward slightly during the insweep in an attempt to 'round off' the
armstroke into the beginning of the recovery. While this may have some advantage in
producing a
Chapter 4 Breaststroke and Butterfly Coaching Swimming: An
Introductory Manual (second edition)
27

28 slightly faster recovery movement, it may also reduce the potential propulsive force
applied during the final stages of the insweep. During all propulsive sweeping actions the
hands are angled so that pressure is felt on the palm surface. Breaststrokers will produce a
powerful upward force component during the insweep; this lifts the shoulders well above
the surface. Lifting the trunk does not cause a problem for streamlining if the hips remain
high in the water. A high hip position determines if the stroke effectively combines
propulsive force with streamlining effects to produce an efficient stroke.

Arm Recovery and Breathing

The recovery phase of the armstroke is combined with a lowering of the head position.
The swimmer usually has the sensation of diving or lunging forward with the arms and
trunk. Hands should remain together as the arms push or lunge forward. Variations are
seen in the position of the palms as the arms extend, some swimmers turn the palms
inward and other swimmers position them facing downward. Both hands may break the
surface of the water during the recovery, but the elbows must remain below the surface to
comply with technical rules. Some swimmers use an over the surface recovery technique
to increase the speed of the arm recovery and thus increase the overall stroke-rate.

Positioning the head so that a breath can be taken during each stroke cycle is an integral
part of timing each stroke. During the insweep of the hands the shoulders will naturally
lift. Taking a breath at this time makes use of this upward component to help stabilize hip
position. If the head is lifted too early in the stroke cycle some of the forward propulsion
will be sacrificed by downward pressure from the arms to support the lifting action. If the
head is lifted too late in the stroke cycle there is no counterbalancing force from the
insweep. As with freestyle and butterfly, there should be a constant exhalation of air from
the nose and mouth whenever the face is below the surface and an increase of exhalation
pressure just before the mouth breaks the surface. Because breaststrokers keep their
shoulders parallel to the surface throughout the stroke (i.e. no body roll) a wave forms in
front of the head and shoulders. The chin must clear this wave during the breathing
action. Some breaststrokers position the head with the chin close to the chest (see photo),
others prefer to keep the face positioned forward (refer to the first photo in this section).
Kick

The leg movement is correctly referred to as a whip kick. It involves simultaneous


extension of the hip and knee and inward rotation of the ankles. From a streamlined body
position at the start of a stroke cycle the recovery of the kick begins with a subtle bend of
the knees as the arms reach the widest point during the stroke (i.e. the transition from
out-to-insweep). As the arms start the insweep the body position changes only slightly;
hip position is held stable as the trunk begins to rises out of the

Chapter 4 Breaststroke and Butterfly Coaching Swimming: An


Introductory Manual (second edition)
29

water. Flexion at the hip and knee lifts the heels closer to the buttocks as the ankles
dorsi-flex (i.e. ankle at a right angle position) and rotate outward. For a split second at the
completion of the leg recovery the body position is poorly streamlined. However, proper
flexion at the hip is necessary to create a powerful thrust during the propulsive phase of
the kick and therefore some resistance against the legs is unavoidable.

Extension of hip and knee joints is made as the arms extend forward and the head begins
to lower. This action positions the trunk, arms, and head so the power of the kick is used
effectively to drive the trunk forward. A slight wave action in the trunk position results
from a combination of actions: (1) lowering the head position to streamline the upper
body; (2) changing from a slight concave shape in the upper back to a straighter or
convex shape (i.e. shoulders rounded forward) as the body dives forward; and, (3) inward
rotation of the feet during the forceful extension of the legs. Leg extension is an
accelerating movement, combined with inward rotation of the ankles these actions helps
to lift the feet slightly at the end of the kick.

1. Feet turn out and legs extend. 2. Ankles rotate inward. 3. Kick complete.
Timing

Timing in the breaststroke is critical to optimal propulsion because any unnecessary


overlap in the application of power impulses will decrease the potential net result. If the
propulsive phase of the kick overlaps too much with the arms there will be a reduction in
net potential propulsive force from the kick. The coach will be able to see when obvious
timing errors are made because the stroke will look uneven. Competitive breaststroke has
changed over the years to reflect more precise timing of stroke components. An ideal
stroke-rate must be determined for each race distance (i.e. 50m to 200m) to take
advantage of one’s stroke capabilities. Current techniques use almost no pause or glide
between the end of one stroke and the start of the next. Breaststrokers using a fast-rate
wave action technique may appear as though the lower body is ‘dolphin kicking’.
However, this is not the case, as a dolphin kick involves downward movement of the feet
with the knees bent. This would not be permitted under the rules for breaststroke
swimming. What’s actually happening is the inward rotation of the feet serves to lift the
soles of the feet at the end of the kick. At the same time the trunk extends, or lunges,
forward to lift the hips. Because the head position is also lowered at the same time, the
affect appears to be an undulating movement of the trunk. Biomechanical analysis shows
us the body’s centre of gravity (located at the hips) does not move up or down too much.
However, the changing position of body segments produces the resulting ‘wave action’
appearance.

Chapter 4 Breaststroke and Butterfly Coaching Swimming: An


Introductory Manual (second edition)

30 Butterfly – Overview

Efficient butterfly technique relies on timing to maximize propulsion while minimizing


resistance. Correct timing of the kick is used to position the body so the large surface area
of the trunk is streamlined at both the least and most propulsive aspect of the armstroke.
The large power impulse generated by the simultaneous double-arm pulling pattern yields
great propulsive potential; however, during the arm recovery phase there is no propulsion
generated. This creates a ‘dead space’ in the stroke. Reducing resistance at key points
during the stroke cycle is a major objective of ideal butterfly technique. The simultaneous
and symmetric arm and leg actions mean the body’s centre of gravity will naturally tend
to rise and fall slightly. Correct timing of the kick and breathing action helps to facilitate
a smooth ‘dolphin-like’, or wave action, that deflects water over and around the body to
help maintain forward momentum.

Streamlining

Before describing the arm and leg action it’s important to note how timing can be used to
help streamline the body. Turbulence is minimised by positioning the body so that water
flows over / under / and around with a gradual, rather than a sharp, deviation from
horizontal flow. The first streamlining technique is to enter the hands at shoulder width,
or slightly closer together as the first kick is initiated. This helps to set the body position
with hips high in the water. At this point the propulsive force generated by the arms is
almost nil; thus, its' important that the hips remain high so water flow along the body
(while not perfectly horizontal) is smooth and unbroken.

1. First streamlining technique is to enter with hands at shoulder width and start the first kick.

A second technique that affects streamlining occurs during the middle stages of the
stroke, when the arms are positioned to deliver maximum drag propulsion. The legs are
stretched toward the surface to level the trunk. This requires strength in abdominal and
lower back muscles, as well as flexibility. This action also helps to counterbalance the
lifting of the head if a breath is taken on that stroke.

2. Second streamlining technique is to stretch the legs up during mid-stroke (prepare for second kick).
Chapter 4 Breaststroke and Butterfly Coaching Swimming: An
Introductory Manual (second edition)
31

A third application of streamlining occurs when the arms are being recovered over the
water. By stretching the legs the soles of the feet come to the surface and the trunk is
positioned closer to horizontal. Streamlining at this point is one way of conserving
forward momentum generated during the powerful finish of the underwater armstroke
cycle and the second kick. Keeping the trunk high in the water allows the arm recovery to
be completed smoothly. The swimmer actually slides forward on top of the wave created
by the body’s forward motion. One of the greatest difficulties encountered by novice
butterflyers is completing a strong second kick. If the second kick is weak, or poorly
timed, the hips will be too low as the hands exit the water. This will encourage the
swimmer to pull the hands sharply upward near the end of the stroke and begin the arm
recovery too soon. Once the timing is delayed, it’s increasingly difficult to lift the soles of
the feet to the surface in time for the start of the next stroke. The end result of this series
of timing and streamlining problems is a rapid deterioration of body position. The
swimmer becomes more vertical in the water and performance suffers.

3. Third streamlining technique is to recover the legs during the arm recovery (prepare for next stroke).

Kick

Dolphin kicking is used exclusively during butterfly swimming, as well as during specific
parts of freestyle and backstroke races. Therefore, it’s a useful crossover skill. Two
kicking cycles fit nicely within each complete butterfly stroke cycle. Although some
swimmers have tried to exist with one kick per stroke cycle, and others have tried to use
three or more kicks, the net result of these off-tempos is usually a reduction in the overall
streamlining of the body. There is some difference of opinion among coaches regarding
the relative contribution of each kick. Some believe the second kick should be stronger
than the first. This belief is strengthened by the observation that over racing distances of
50m and 100m, swimmers tend to accentuate the second kick as a result of the great
acceleration in hand speed during the second half of the armstroke. Other coaches feel
that each of the two kicks should be the same depth and intensity; this is the case for most
200m swimmers. There is general agreement that all butterfly swimmers should attempt
to get the most out of each kick by timing the downbeat precisely at certain phases of the
armstroke.

The dolphin kick begins with the legs together and extended, toes pointed. Extension at
the hip (or hyperextension) lifts the feet to the surface. The knees begin to flex while the
heels stay at the surface. Knee bend rapidly increases until the angle between the
hamstrings and calf muscles is almost 90-degrees. The propulsive thrust of the legs is
accomplished by forceful extension of the knees and simultaneous flexion of the hips to
create a ‘whip like’ downward movement of the legs. During this propulsive thrust the
abdominal muscles must remain tight to stabilise the pelvis

Chapter 4 Breaststroke and Butterfly Coaching Swimming: An


Introductory Manual (second edition)

32 and allow the large muscles of the legs and buttocks to contract strongly. The
backward-downward push from the legs will drive the hips upward and forward. When
the downward thrust of the kick is completed the hips are still flexed slightly.

Armstroke – Propulsive Phase

The pattern traced by the hands (relative to the body) during butterfly is sometimes
referred to as an ‘hour-glass’ shape. There are three or four sequences of sweeping
movements that contribute to propulsion in the butterfly. Hands enter the water at
shoulder width, or slightly inside, palms pitched outward so the thumb edge of the hand
enters first. Hand position is angled upon entry to allow air bubbles to escape the pulling
surface. The angled hand entry also allows a smooth transfer of momentum developed
during the arm recovery as the hands strike the surface. The first downbeat of the kick is
completed almost immediately as the hands enter and begin to press outward. Hands
sweep symmetrically outward with the elbows held up. Some amount of lift is generated
at this point of the stroke; however, the major propulsive impulse will not come until the
hands move under the body. The insweep may also be angled slightly up toward the body;
this determines how much elbow- bend is achieved. A final propulsive sweep, or push,
moves the hands from a position under the chest to a position past the hips. The hands
slide out of the water and into the recovery phase in one smooth motion.

Sweeping movements cause the hands to change direction several times (i.e. from
'out-and-down' to 'in’ and then 'back-and-out'). Because the movements are rounded,
rather than changing direction sharply, hand speed continues to accelerate during the
length of the stroke. There is a large amount of drag propulsion generated during the
butterfly armstroke because the surface of the hand-forearm can be positioned at right
angles to the directional movement of the body.

Armstroke – Recovery Phase

When the underwater push of the hands is completed, the wrists relax to turn the palms
inward. The recovery phase begins as the elbows begin to lift and the hands slide out of
the water with the little finger edge of the hand leading. At this point the downbeat of the
second kick assists by positioning the hips close to the surface. The arms remain
comfortably extended as they begin to swing forward. Once the hands lift sufficiently to
clear the water the forward swing of the arms becomes a ballistic movement. Lateral
forces created by the wide swing of right and left arms during the recovery act to cancel
each other; this keeps the shoulder position stable.

Chapter 4 Breaststroke and Butterfly Coaching Swimming: An


Introductory Manual (second edition)
33

Breathing

Most butterfly swimmers breathe on every second stroke cycle. If the stroke-rate is low,
or the swimmer needs to breathe more frequently because of the race distance (i.e. 200m
event), taking a breath every stroke cycle may be required. Swimmers who develop
correct timing and have a strong kick should not worry about reducing their stroke
efficiency if they breathe during every stroke cycle. Conversely, during sprint events of
50-100m the swimmer may want to limit breathing for 3 or more strokes to achieve a
faster stroke-rate. The breath is always taken in conjunction with the natural rise of the
shoulders created by the strong propulsive force generated during the middle of the
armstroke. The head is generally lifted with the face positioned forward, chin just clear of
the bow-wave. However, some swimmers may prefer to turn the head to the side, similar
to the movements used in freestyle. Side breathing is much harder in butterfly than in
freestyle because the shoulders do not roll; the action is truly a turning of the neck. Side
breathing may be an advantage to some swimmers because it doesn't require as much
lifting of the head and shoulders.

Timing

When the face is below the surface the line of sight is down and slightly forward. As the
hands sweep inward and backward there is a natural lifting of the shoulders and the chin
should begin to move forward. Before the arms finish their propulsive phase, the face
breaks the surface. Timing of the second kick is critical because it helps to hold the head
up while the breath is completed and the arm recovery begins. As mentioned, the second
kick also maintains trunk position at a time when the final push of the arms deliver peak
propulsive force. As the arms swing forward the head is lowered so the face is in the
water before the next hand entry is made. This action of dropping the head serves two
purposes: (1) it allows the muscles of the upper back to complete the forward swing of
the arms without restriction, and (2) it lowers the body's centre of gravity to assist the
wave action. Swimmers who keep their head up too long will eventually disrupt their
body position; the legs will drop and greater frontal resistance will result.

As we have seen, correct butterfly technique is all about timing. Each stroke contains
periods of accelerating movement that must be performed simultaneously on right and
left sides of the body. Unskilled swimmers often fail to coordinate the symmetry of the
stroke. The timing of each kick is also critical. Both kicks help to maintain forward
momentum and body position. Timing of the head movement must be precise so that
body position is not adversely affected. The breath must always be taken in coordination
with the stroke, without breaking stroke rhythm. When the components of the stroke are
properly coordinated, the body's centre of gravity should rise and fall in a narrow wave
pattern.
Chapter 4 Breaststroke and Butterfly Coaching Swimming: An
Introductory Manual (second edition)
34

CHAPTER 5 STARTS, TURNS, and


FINISHES

Dive Start

Although a number of techniques have been used in the past, the simplest and most
effective racing dive is called a ‘grab start’. A common variation of the grab start is the
‘track start’ because it resembles the starting position used in sprint running. The
difference between the two techniques is the initial placement of the feet on the starting
block; this affects the forward transfer of bodyweight. One start is not necessarily better
than the other, it’s a matter of individual preference and how the start is executed. Either
start can be used for freestyle, butterfly, breaststroke and individual medley races.

A long-whistle from the referee signals the competing swimmers to step onto the block
and assume any position they desire. The swimmer may take up a fixed starting position
straight-away or stand back from the front-edge of the block and await further starting
instructions.

On the starter's command "take your marks" all swimmers must immediately take up a
starting position with at least one foot at the front edge of the block. Whatever starting
position is taken, the swimmer’s bodyweight should be balanced over both feet to
maintain a stable base of support and allow the swimmer to remain motionless until the
starting signal (i.e. usually a gunshot, horn or beep) is given. Foot placement (width) may
be close together or slightly apart. Hand placement may be next to the feet at the front
edge of the block or grasping the side edge of the platform. Hand and foot placement
should be comfortable and allow for some pressure to be maintained against the surface
of the block, this helps to maintain stability. The swimmer will be in a position where the
hips and knees are flexed. The degree of flexion will depend upon a swimmer’s
individual flexibility, balance, and foot placement. The head is positioned slightly upward
so the water is in view.

There are several objectives to be achieved during the dive. First, the swimmer must react
quickly and move bodyweight forward after the starting signal. Second, the trunk and
legs must be positioned so that maximum force can be applied against the block. Third,
during the flight phase the

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)
35

body must be repositioned so that entry into the water is made with minimal resistance.

The first objective, reaction time, will improve if the swimmer learns to concentrate on
the starting signal, rather than be distracted by other external cues. It’s also an advantage
to maintain some amount of muscle tension when the hands are in contact with the block.
The second objective is realised if the swimmer can move the trunk forward to a position
that is approximately horizontal when the legs reach full extension. The first photo shows
a swimmer in the start position, the pathway that the body’s centre of gravity (COG) will
take during the flight phase has been plotted (i.e. connected dots). The COG initially
drops slightly when the trunk begins to move forward and will follow a parabolic
pathway once the feet break contact with the starting block. The distance the COG travels
during the flight phase is established by the accumulated muscular force transferred
through the segments of the lower body. Also, when the feet leave the block the axis of
the trunk is angled slightly upward from horizontal (solid lines in the photos). To achieve
the third objective of a good dive the swimmer must allow the axis of the trunk to rotate
(i.e. tilt downward) so it lines-up with the pathway of the COG. This adjustment in trunk
alignment is accomplished by performing several small actions that serve to redistribute
the weight of the body around the COG. The easiest way to bring weight closer to the
COG is by dropping the head between the arms during the flight phase. Another way of
adjusting body position is to bend slightly at the knees or hip; once again, to distribute
more bodyweight forward. The net affect of these actions will be to rotate the trunk
downward to a position that’s in line with the flight pathway (third photo). The entry is
made with the head positioned between the extended arms and the hands held together,
punching a small ‘hole’ in the water and allowing the body to smoothly follow through
the hole along the pathway of the COG.

Dive Start Checklist


• Feet firmly on the block, bodyweight balanced.
• Knees and hips bent.
• Looking down and slightly forward.
• Hands in contact with the block, apply slight amount of pressure.
• Bodyweight begins to move forward, arms in front of body, head up.
• Arms push or swing forward, but stop below the shoulder, head up as feet leave the block.

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)

36 • Full leg extension off the block.


• Head moves down between the arms during flight phase.
• Slight bend at the hip or knees to help tilt the trunk downward during flight phase.
• Reach for the water with hands together, head between the arms.
• Enter the water by making a small hole with the hands.
• Keep body extended upon entry, but allow the legs to 'relax' slightly as they enter.
• Point the toes on entry.

Backstroke Start

Backstroke events (and the backstroke leg that starts a Medley Relay) begin from
water-level. This takes away the advantage of an elevated starting position. However, the
principles of a good dive start also apply to the backstroke start. The swimmer takes up a
position with the feet fixed firmly on the wall, at about hip width apart. Some swimmers
feel more comfortable with one foot slightly lower; this increases the size of the base of
support. Hands may grip the starting block in several different positions, depending upon
block design. The most common hand positions are directly in front of the body (grip
with palm facing downward) or to the side of the block (palms facing inward).

On the long whistle the swimmer positions the hands and feet. Remember that the toes
must remain below the water surface. At the command "take your mark" the swimmer
pulls the body slightly upward The hips should be as high as possible, but not so high that
the swimmers looses balance and foot pressure against the wall surface.
When the starting signal is given the swimmer must try to drive the hips upward and
backward while throwing the arms over the head. Several simultaneous actions must be
completed in a split second. Extension of the hips must initiate the movement, but a
forceful swing of the arms (either to the side or straight over the trunk) will help to
project the body forward. The head tilts back and the trunk is arched slightly as these
movements are made. The hands meet when the arms reach full extension and the final
drive from the knees and ankles completes the movement. The hips should be clear of the
water surface, legs together and the toes pointed. As the hands enter the water a slight
flexion at the hip and knees are used to lift the legs and allow the body to slide into the
water smoothly (see photos that follow).

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)
37

1. Hand entry with arms together. 2. Flip feet up to straighten body for entry.

Once completely underwater it's possible to level the body by using the hands as a rudder.
The upper body must be held in a streamlined position with arms extended and head
firmly between the upper arms. Dolphin kicking is used to propel the swimmer forward
while submerged. The swimmer may need to exhale continuously through the mouth and
nose to keep water clear of the nose while underwater. As the swimmer approaches the
surface, within the 15m limit for underwater swimming allowed by the rules, the kicking
technique changes from dolphin to flutter as the first armstroke begins. The face should
break the surface during this first armstroke. If the timing is correct the swimmer will
maintain momentum and should continue into rhythmic stroking at race pace.

Backstroke Start Checklist


• Feet firmly on the wall (usually slightly apart) so bodyweight is evenly balanced. Toes
must remain below the surface. Feet may be placed parallel or staggered (i.e. one foot
slightly lower than the other) on the wall.
• Knees bent, bodyweight close to the wall.
• Line of sight is forward.
• Hands grip either the horizontal or vertical bar in a comfortable position.
• On the starting signal push or throw the arms back.
• Slight bend in the back to help lift the hips.
• Reach for the water with hands together, head between the arms.
• Try to ‘flip’ the feet upward (by bending at the knees) as the hands enter.
• Enter the water with body straight, moving through a small hole made by the hands.
• Keep trunk extended upon entry, but allow the legs to relax slightly as they enter.
• Point the toes on entry.
• Dolphin kick several times in a streamline position on your back while
underwater. Head must break the surface within 15m of the start.
• Begin normal kicking action as the first armstroke begins.

Racing Turns (Overview)

The tumble-turn technique used in freestyle is essentially the same for backstroke. Even
young competitive swimmers, from the age of 8 years, are capable of mastering the skills
of this turn. Tumble-turns should be an integral part of all training swims. The energy
cost (provided the technique is correct) is not significantly greater than the

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)

38 slower hand-touch turn. The major problem encountered by young or inexperienced


swimmers in executing tumble-turns is insufficient forward momentum going into the
turn. This is usually the result of: (1) poor streamlining which results from lifting the head
out of the water to look for the turn or take a breath, or (2) reducing or stopping the flutter
kick during the pre-turn stroke, or (3) not using the arms effectively to maintain
momentum at the start of the tumbling action.

Breaststroke and butterfly turns are similar in many ways because these two strokes are
swum with the shoulders level (i.e. no roll of the shoulders as in freestyle and
backstroke). The turn must pivot the body after both hands have touched the wall. The
pivot action turns the body sideways so that wave action against the trunk is minimised.
The turning action must then return the swimmer to a level shoulder position during the
push-off to continue the race. A fast competition turn in breaststroke or butterfly uses the
body's forward momentum as the hands touch to begin the turning action. The legs are
drawn up into a tuck position so the feet can be quickly positioned on the wall for the
push-off. During the turn one hand stays below the surface to help the body pivot in a
tuck position and the other arm comes over the surface to speed the turning action of the
shoulders. Both hands meet in front of the body as the arms extend and the legs drive the
body off the wall in a streamline position.

Freestyle Tumble Turn

1. Approaching the turn. 2. Last armstroke completed.

3. Head down, dolphin kick. 4. Tuck position during rotation.

5. Feet strike the target, hips are flexed. 6. Extend upper body, drive with legs.

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)
39
7. Streamline body off the wall. 8. Several fast dolphin kicks may be used.

Freestyle Turn Checklist


• Complete last armstroke so that both hands remain near the side of the body.
• Drop head and dolphin kick.
• Tight tuck at the hips, keep the chin close to the chest.
• Bend knees as the legs rotate out of the water directly over the body.
• Use both hands to help the body rotate while it's upside-down.
• Both feet strike the wall at the same time, body is in a tuck position, hands close to the
face.
• Extend forward with the arms and push through the hips and knees to stretch the body
completely during the push-off.
• Head between arms, hands together during push-off.
• Fast, sharp dolphin kicks may be used (keep upper body streamlined).
• Begin first stroke and strong transition to freestyle kick.

Backstroke Tumble Turn

1. Approach the turn, count strokes. 2. Recovery-arm crosses over the body.

3. Dolphin kick into tuck position. 4. Rotate in tuck position.


5. Push-off, body streamlined. 6. Dolphin kick underwater.

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)

40 Backstroke Turn Checklist


• On the last stroke use recovery-arm to begin body rota tion by crossing-over the body for hand
entry in front of the opposite shoulder
.
• Finish the underwater pull as the head drops down.
• Small dolphin kick to lift the hips as the tumble be
gins.
• Rotate in a tuck position to complete the tumble.
• Keep both hands und
er the face as the body tumbles, this allows the arms to assist the body's rotation.
• Feet strike the wall at the same time with the toe
s pointed up.
• Push-off by extending from the hips and knees.
• Stretch the body
completely in a streamlined position, head between the arms and hands together.
• Dolphin kick in a streamlined position, stretch upward wi
th the arms.
• Begin first stroke with a strong transition to flutter kick.

B reaststroke Turn
1. Time the last stroke into the wall. 2. Touch at full extension (palms flat).
3. Legs tuck and hips rotate toward the wall. 4. Underwater hand helps to level the trunk.
5. Push-off in a streamline position. 6. Streamline glide in prone position.
7. Long underwater armstroke. 8. Finish pull, stay in streamline position.
Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An Introductory Manual (second edition)
41
9. Recover arms close to the trunk. 10. Kick to drive the head to the surface.
Breaststroke Turn Checklist
• Reach for the wall and begin to pull the knees up under the body.
• After a two-hand touch, pull one hand away (this arm stays below the surface).
• Body pi vots toward the leading arm (this arm applies pressure to help rotate the body).
• Trailing arm falls off the wall and recovers over the wa
ter.
• Both feet strike the wall with toes pointed to the side.
• The arm that is out of the water travels directly over
the head and enters in front of the body (i.e. to meet up with the underwater arm).
• Body is level as extension from the hips and knees drives the push-off.
• Body quic
kly rotates from a slight side position into a prone (i.e. stomach down) position.
• Stretch and streamline the body with head between the arms and h
ands together.
• Underwater pull is similar to a butterfly pull, keep the head level.
• Underwater pull finishes with both hands at the side of the bod
y.
• Hands recover close to the body as the legs begin to recover.
• Legs drive back (breaststroke kick) as arms stretch forward.
• Head lifts and breaks the surface as the first armstro
ke is made (head must break the surface before the hands start to move inward).
B utterfly Turn

1. Contact the wall on a full stroke. 2. Pull knees under the body (one hand under).
3. Pivot body onto its side, feet against the wall. 4. Push-off in a streamline position.
Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An Introductory Manual (second edition)

42 Butterfly Turn Checklist


• Stretch as you reach for the wall, hand touch on a full stroke.
• Pull the knees under the body as a two-hand touch is made.
• Leading arm m
oves underwater and trailing arm moves over (similar to Breaststroke).
• Use pressure on the hand of the leading arm to help the body rotate.
• Trailing arm comes directly over the head and enters in front of the body.
• Hands together as the hips and knees extend to drive the body off the wall.
• Body com
es off the wall slightly on the side, but quickly rotates to a prone position.
• Point the toes during the streamlined glide.
• Several dolphin kicks, keep the trunk level and head between the arms.
• Begin first butterfly armstroke as head breaks the surfac
e (within 15m limit).
• Try to finish the first stroke with the face in the water.

In
dividual Medley Turns
Finally, the Individual Medley event has its own set of turns to master as the swimmer switches
from one stroke to another. These turns are specialised and will be taught once individual turns in
each of the four strokes are mastered. Swimming rules require the swimmer to complete each
section of the race in accordance with the specified rules for that stroke
, and begin the next section of the race in accordance w ith the rules for that stroke.

The butterfly to backstroke changeover begins as a normal butterfly touch on the wall. The
swimmer should try to make contact with the wall on a full stroke, with arms extended and
without excessive gliding into the wall. Once a simultaneous two-hand touch is made, the
swimmer will pull one hand away (i.e. leading-hand) and begin to pivot the legs toward the wall
in a tuck position. The trunk begins to twist slightly during the pivoting action of swinging the
legs up. When the feet strike the wall the toes will be pointing straight up. The momentum of the
body’s rotation causes the trailing-hand to fall off the wall an instant before both feet make
contact. The trailing-hand is usually brought over the surface and re-enters the water behind the
head as both arms extend. A streamlined position on the back is held for only a fraction of a
second; then underwater dolphin kicking (in a supine position) is used to drive the
body toward the surface. The face breaks the surface as the first armstroke b
egins.
1. After two-hand touch, pivot trunk. 2. One had over (one under) to level the body.
Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An Introductory Manual (second edition)
43

The backstroke to breaststroke changeover is perhaps the most difficult of the medley
turns. The backstroke leg must end with a hand touch while on the back. There are
several methods commonly used to execute this turn. Beginning competitive swimmers
will find that a simple open-pivot turn is effective and then graduate to a faster method.
The hand touch is always made at, or slightly below, the surface. In the open-pivot turn
the body pivots underwater, rotating toward the hand making the touch, as the hips and
knees flex to bring the body into a tuck position. The non-touch hand remains underwater
and assists with the rotation of the body by pressing upward with the palm. The shoulders
and head move toward the underwater hand as the touch-hand releases from the wall just
before the feet strike. The toes will point to the side. The touch-hand comes over the
water and enters in front of the head; both arms then extend as the push-off is completed.
The body may be slightly on its side as the legs begin to extend, but the rotation of the
turn quickly brings the body into a prone position during the glide off the wall. The long
underwater breaststroke pull completes the transition to the next leg of the race. Once a
swimmer has mastered the open-pivot turn it’s time to move on to something faster. A
second stage turn usually involves lifting the feet out of the water during the pivot.
Because most of the legs are lifted out of the water during this turn there is less water
resistance and it should be faster.

1. Hand touch (shoulders not past 90 degrees). 2. Swing legs under body (take a breath).

3. Hand over the water. 4. Streamline body position during push-off.

A breaststroke to freestyle changeover is very simple. The touch is made from the
breaststroke and the pivot action is identical to a normal breaststroke turn. As the
swimmer drives off the wall a normal freestyle kicking action begins. The success of this
changeover is usually determined by the swimmer's ability to pivot quickly and change
swimming tempo quickly (remember, the swimmer is changing from the slowest to the
fastest stroke). Chapter 5 Starts, Turns, and Finishes
Coaching Swimming: An Introductory Manual (second edition)
44 1. Touch on a full stroke (arms extended). 2. Swing feet under the body (tuck position).
3. Swing arm over the water (take a breath). 4. Extend body for streamlined push-off.

Race Finishes

Finishing a race correctly can save time and often provide the winning margin in a close
contest. There are several simple techniques any swimmer, from beginner to elite, can
execute. In all strokes it’s important to practice finishing on a full stroke with the arm(s)
extended; this should be practiced daily. Swimmers can learn to adjust stroke length
slightly while still 5+ metres from the wall. Large adjustments to stoke length should not
be necessary on the very last stroke. In strokes where a one- hand touch is permitted (i.e.
freestyle and backstroke) it’s possible to hyperextend at the shoulder by rotating the trunk
slightly onto the side as the final reach for the wall is made. However, backstrokers must
take care that the shoulders do not rotate past 90 degrees before the hand contacts the
wall, so that a legal touch is made while still on the back. In all finishes the head should
stay down as the final stroke is made. There should always be sufficient forward pressure
from the touch to allow any electronic timing equipment to activate. Swimmers who try
to touch as they lift their head and pivot to look at the electronic timing display or look up
into the grandstand will often miss the wall or slide across the touchpad. Another time
saving technique on freestyle and butterfly is to limit breathing during the last five metres
of the race. This allows the swimmer to visually focus on the wall for the final lunge.
Backstrokers must learn to count their strokes from flags to wall and know exactly when
to lunge backward with one arm to make their final touch.

Finishing Checklist
• On freestyle, breaststroke and butterfly finishes look for the wall targets and judge the
distance travelled over the last 3-4 strokes. This should be practiced in every training
session.
• On backstroke count strokes as the head passes under the backstroke flags. This must be
practiced at race pace.

Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An


Introductory Manual (second edition)
45

• Once you know the number of strokes taken from flags to wall (this applies to Fly –
Back – Breast – Freestyle strokes), try to finish with a full reach on the final stroke.
• On freestyle and backstroke finishes, as the arm reaches for the wall the body may
rotate slightly on its’ side (be careful that the shoulders do not rotate past 90 degrees
before the touch is made in backstroke). This helps the swimmer extend the distance of
the final reach toward the wall.
• Keep the head down until the hand(s) touch the wall (all strokes).
• Keep the body streamlined until the touch is made. This is done by using small, fast
kicks when swimming freestyle, backstroke, and butterfly.
• Small adjustments to stroke length during a breaststroke finish are made with the
armstroke, not with the kick.
• On freestyle and butterfly, try to finish the last 5m, from flags to the wall, without
taking a breath.
Chapter 5 Starts, Turns, and Finishes Coaching Swimming: An
Introductory Manual (second edition)
46

CHAPTER 6 PHYSIOLOGICAL
CONSIDERATIONS
Competitive swimming success is dependent upon many factors. The application of
mechanically efficient stroke technique has already been discussed; there are also certain
physiological demands associated with swimming, in particular swimming at race speeds.
Energy supply capabilities and muscular capacities must be trained to meet these
requirements. All swimming coaches must continue to update their knowledge of the
body's physiological response to exercise. Only then will the coach have confidence in
planning the best possible training program. As a beginning coaching text, this chapter
will not attempt to explain exercise physiology in detail. However, the coach will be
introduced to a number of principles. The application of these principles will be
fundamental to the development of a sound training program.

Energy Supply for Swimming

Muscle contractions require an energy source. In the broadest sense, our energy supply
comes from the foods that we consume. However, a number of processes must occur
before food sources are broken down into compounds useable at the cellular level. The
basic energy source required by cells is a compound called adenosine triphosphate (ATP).
Each ATP molecule is structured so that one adenosine component is bonded with three
phosphate components. The bonds represent stored energy. The presence of muscle
enzymes causes one of the phosphate components to separate, thus releasing stored
energy when the bond is broken. In addition to the release of energy we now have the
molecule adenosine diphosphate (ADP) containing only two phosphate components. Our
continuous energy demands are satisfied by the different systems our body uses to break
down molecules to release energy and then reform the energy source. There are three
systems that contribute to the total supply of energy:
• ATP-PCr (called the phosphagen system),
• anaerobic glycolysis (called the lactic acid system),
• aerobic system (called the oxygen system).

Each energy supply system has distinct features that determine the rate at which ATP
stores can be used and then regenerated. The rate of energy supply is determined by the
energy demand required to perform a movement task. Thus, to swim 50m successfully in
competition requires a very high rate of work (i.e. energy demand) and to swim 1500m
requires a lower rate of work, but sustained over a longer period of time. Understanding
the energy supply mechanisms will influence how the coach plans training as well as how
the coach plans race strategy.
The ATP-PCr system uses the presence of another phosphate compound called
phosphocreatin, or PCr, to recombine with ADP and produce ATP. The system actually
rebuilds the existing structures so that high-energy compounds are available 'on site'
within the muscle cell for immediate energy release. The breakdown of PCr releases the
energy required to drive this chemical reaction. The process can replace ATP so quickly
that consecutive muscle contractions may take place rapidly. Unfortunately, the supply of
PCr in the muscle is limited. When rapid energy demand

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)
47

calls upon the ATP-PCr system, it can supply the major portion of energy requirements
for only a few seconds. These chemical reactions do not rely on the presence of oxygen,
and therefore are classified as anaerobic metabolism.

Anaerobic glycolysis is another energy system that does not require oxygen. Glycolysis is
the metabolic process of breaking down the simple sugar glycogen. The series of
chemical reactions (i.e. glycolysis) resynthesizes glycogen to produce ATP molecules.
This energy pathway reacts quickly when the energy demand is high and can be sustained
at a very high level for about 40-50 seconds. However, the breakdown of glycogen is
incomplete and lactic acid is produced as a by-product. When lactic acid accumulates in
the muscle tissue it changes the chemical environment and this becomes a limiting factor
to sustained high energy supply. Accumulation of lactic acid in the muscle tissue
increases the concentration of hydrogen ions (i.e. positively charged hydrogen atoms) and
this makes the chemical balance more acid. The characteristic muscle burning sensation
and tightness experienced after a few seconds of high intensity exercise is the signal that
chemical reactions are slowing down and energy from anaerobic glycolysis has reached a
peak.

The aerobic system relies upon a series of chemical reactions to break down more
complex fuel sources into ATP, the system relies upon the presence of oxygen to help
break down molecules and release energy. There are hundreds of chemical reactions
required, making it a more complex system than anaerobic glycolysis. This means the
aerobic system takes a little longer to reach peak energy supply. The system is also
limited by the body's ability to deliver oxygen to the muscle. However, there are a
number of advantages to obtaining energy supply from the aerobic system, rather than
from anaerobic processes. First, under aerobic conditions both fatty acids (i.e. simple fat
molecules that are transported in the blood) as well as glycogen can be utilized, giving a
greater fuel storage potential. Protein sources may also be used, but under exercise
conditions this complex breakdown is usually not required. Second, the synthesis of ATP

is more efficient, the by-products produced are carbon dioxide (CO2) and water (H2O).

Carbon dioxide is easily removed from the cell, transported via the bloodstream and
expired through the lungs. Excess water is also removed from the cell easily without
major changes to the cell chemistry.

Interaction of the Energy Systems

Both anaerobic and aerobic metabolic pathways contribute to the total energy production
during all levels of activity, from the shortest sprint to the longest distance swim.
However, the relative contributions of the three energy systems are dependent upon the
energy requirement in terms of 'how much' and for 'how long'. Because competition
swimming involves the application of peak or optimal resources, we generally require
very high amounts of energy over a short period of time during sprint swims, and lower
amounts of energy over a longer period of time during distance swims. Race distances of
100-200m fall between the two extremes and require different proportions of energy
resources from the anaerobic and aerobic systems.

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)
48

Relative Contribution of Anaerobic and Aerobic Energy Sources

Time Percent of Supply Percent of Supply


Anaerobic Sources Aerobic Source

10 sec. 95+ <5 30 sec. 85 15 1 min. 75 25 2 min. 60


40 4 min. 40 60 9 min. 20 80 15 min. 10 90

We determine the relative energy contributions from aerobic and anaerobic sources by
monitoring two key markers: (1) oxygen consumption, and (2) blood lactic acid
accumulation. We do not measure oxygen consumption directly during swimming unless
laboratory conditions exist. However, we can observe and measure respiration rate and
heart-rate to get an indication of increased oxygen delivery. Lactic acid measurement can
be taken from the blood to give us an indication of the chemical balance in the muscle.
Most coaches will not have access to blood lactate measurement. However, we can get
some indication of the anaerobic energy contribution by monitoring perceived effort (i.e.
subjective evaluation of effort) and the relationship between work-rate and swimming
velocity.

At rest our energy demands are low, so low-level aerobic metabolism is an efficient way
of supplying energy. Any amount of lactic acid produced (remember that both aerobic
and anaerobic pathways are always operating) in the muscle is efficiently removed or
metabolized. Thus the muscle pH remains unchanged and there is no sensation of muscle
discomfort. Lactic acid measured in the blood during rest usually remains around 1
millimole per litre of blood (i.e. mMol/L). A mole is a given amount of a chemical
compound by weight, as determined by its molecular structure.

When energy demand increases, both aerobic and anaerobic systems will respond.
Anaerobic energy pathways can respond very quickly to a demand for high amounts of
energy because increased oxygen delivery is not a factor. Increased oxygen delivery to
the muscle may lag behind at the start of intense exercise. Swimmers having a large
aerobic capacity will be able to gear-up more quickly. If the energy demand remains high
for more than a few seconds lactic acid production begins to overcome the rate of lactic
acid removal from the muscle tissue. Alkaline substances and proteins within the muscle
(these substances are called muscle buffers) will act to absorb some of the hydrogen ions,
but they simply can not keep up. The increased blood supply to the muscle triggered by
the increased demand for aerobic energy production will also help in the lactic acid
removal process. However, the mechanisms of lactic acid removal will not be able to
keep pace with lactic acid production and the chemical balance of the cell will change. If
demand for high amounts of energy continues, the peak rate of anaerobic metabolism will
reach its limits.

The rate of anaerobic energy production will drop back if sustained high energy demand
continues. If the aerobic system has the capacity to supply greater amounts

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)
49
of energy the swimmer will be able to sustain a swimming velocity that represents a
'critical point' where lactic acid accumulation and reduction are balanced (this is often
referred to as the anaerobic threshold or AT). Every swimmer has an individual anaerobic
threshold, represented by a specific swimming velocity and lactate accumulation level
(i.e. number of mMol/L of lactic acid in the blood) that can be sustained for several
minutes. Aerobic fitness is associated with the velocity at which AT is reached. Training
helps to improve the swimming velocity (i.e. making it faster) at anaerobic threshold.
Training will also increase the length of time a swimmer can maintain AT pace and
counteract the affects of lactic acid accumulation. In addition to relating training
adaptations to AT velocity, the AT velocity can be compared to the percentage of total
aerobic capacity it represents. Unfit swimmers who try to sustain fast swimming will
reach their AT quickly (i.e. a fairly low percentage of their aerobic potential) and
gradually begin to slow down as they ‘tighten up’. Fit swimmers have adapted their
energy supply systems so they reach anaerobic threshold at higher swimming velocity
(i.e. representing high percentage of their aerobic capacity) and sustain that speed for
longer periods of time.

Energy System Capacities

Each swimmer's maximal capacities change as a direct result of training, but they're also
influenced by a number of other factors. First, genetic potential influences everything.
Some swimmers will have a natural advantage; this does not always guarantee success,
only a head start. Second, growth and development strongly influence both aerobic and
anaerobic capacity. Physical size alone; with respect to body mass and lung capacity; has
a great influence on one's potential aerobic capacity. The influence of biological maturity
on hormonal activity and muscle development also helps to determine anaerobic
potential. Early maturing swimmers have a temporary performance advantage; however,
in the long-term swimmers who mature later will catch-up and may in fact experience
some long-term performance advantages. Third, mechanical proficiency (i.e. the ability to
move through the water with minimum resistance) can never be ignored. Having large
energy capacities means little if that energy is wasted.

Having the largest possible capacity to deliver oxygen (i.e. max VO2) to the muscles will
contribute to swimming success, regardless of whether a swimmer has a sprint or

endurance orientation. Changes in max VO2 occur during childhood and puberty as a
combined result of maturation and training. Once a swimmer has passed puberty
increases in aerobic capacity tend to be less from year to year. However, mature
swimmers incorporating an appropriate amount of endurance based training program can

still stimulate an increase in max VO2 up to their genetic potential. Endurance


performance will also improve as a result of changes to the swimmer's individual

anaerobic threshold, as a percentage of max VO2. Changes occur to the oxidative capacity
of muscle fibres at a cellular level. Training overloads produce adaptation to a swimmer’s
aerobic capacity and then maintaining a new level of fitness requires relatively less
training volume. This is sometimes called maintenance level training. Aerobic capacity
will decline when the aerobic component of training falls below maintenance level; this
fact has training implications during the 'taper' period and transition periods between
seasons.

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)

50 Peak anaerobic capacity is dependent upon anaerobic glycolysis, which in turn is


dependent upon the supply of glycogen stored in the muscle and the type of muscle fibres
having a glycolytic capacity. Not all muscle fibres are alike. Skeletal muscle consists of
both slow-twitch (ST) and fast-twitch (FT) fibres, each having different capacities.
Slow-twitch muscle fibres have these properties: (1) a low neural activation level and
relatively slow contractile speed, (2) a high oxidative capacity and large blood supply,
and (3) a high resistance to fatigue. Fast-twitch muscle fibres have these properties: (1) a
high neural activation level and fast contractile speed, (2) a high glycolytic capacity and
no direct blood supply, and (3) a low resistance to fatigue. In general, we associate ST
fibres with aerobic endurance capacity and FT fibres with sprint capacity. The percentage
of ST and FT fibres making up the skeletal muscle is genetically determined. Some
people would say that talented sprinters or distance swimmers are born to be that way.
However, it's not as simple as that; the type of training program used will have a great
deal of influence on how some of the FT fibres respond.

Studies have shown that an average person will have an ST – FT distribution in the range
of 40-60%; that is, somewhere between 40% ST / 60% FT and 40% FT / 60% ST.
Therefore, the vast majority of swimmers will have the potential to swim well at most
distances and a few very exceptional swimmers will have greater potential for either
sprint or endurance events. Fast-twitch fibres are actually divided into two sub- groups
(note: some scientists would argue there are more than two sub-groups), known as FTa
and FTb. Fast-twitch, type 'a' fibres tend to alter their functional characteristics based

upon the way they are used over a period of time. This is because FTa fibres have

characteristics that support oxidative capacity of the total muscle and have a greater

resistance to fatigue than FTb fibres. We might say these 'a' type fibres can learn to
function more like ST fibres while retaining many of their fast contractile properties.
Specificity of training allows these fibres to adapt and produce the sustained speed
required for 200-400m events. It also means that training programs that are based
exclusively on sprint work will train the 'a' fibres to function more like the 'b' fibres; this
produces the 'drop-dead sprinter'.

The capacity to use high amounts of energy to produce fast, and strong, muscle
contractions is also dependent upon the neuromuscular recruitment of fibres. When a
submaximal amount of muscle tension is required for a sustained period, ST fibres are the
first ones recruited for the job. If maximal muscle tension is required for sprinting, or
greater muscle tension is required at slower swimming speeds (because of technical
inefficiencies or increased resistance loads), additional fibres must be recruited. The rate
of muscle fibre recruitment will influence energy demand.

Glycogen storage capacity of the muscle is another variable that affects potential
performance. Muscle glycogen is the sole source of ATP synthesis for anaerobic
glycolysis and glycogen reserves are limited. Muscle glycogen level will be dependent
upon diet and the rate at which the available glycogen is used. The individual fibres most
frequently recruited during exercise will be the ones depleted first; thus both ST and FT
will be affected. When ST fibres are depleted of muscle glycogen it's still possible for
them to metabolise fats aerobically. Since FT fibres are reliant on glycogen breakdown
alone to supply energy, depletion of muscle glycogen will limit a swimmer's ability to use
FT muscle fibres to swim fast. The characteristic

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)
51

sensation of muscle fatigue and heaviness is often the result of insufficient energy
reserves in the muscle.
The capacity for PCr energy production is limited by the amount of phosphocreatine
available in the muscle, which is subject to rapid depletion. There is some evidence that
diet may be a factor influencing the amount of phosphocreatine in the muscle, although
more research must be done in this area. The overall potential of this energy system

seems to be determined by the percentage of pure fast-twitch (FTb) muscle fibres.


Phosphocreatine is depleted very quickly during intense activity; however, the
regeneration rate is also very fast. Replenishment of phosphocreatine will take place

under conditions of complete rest or low level activity that does not require the FTb fibres

to fire for a short period of time (perhaps 46-60 seconds). The overall reformation of
phosphocreatine will gradually reduce when repeated short sprints are performed.

Cardiovascular Considerations

The circulatory and respiratory systems contribute to the energy supply process by
delivering oxygen and fuels and removing waste products from the working muscles. The
heart and blood vessels respond to training by becoming more efficient in their delivery/
removal function. Not only does the heart respond to exercise by increasing the rate of
contraction (i.e. heart-rate), but over time the heart adapts by being able to pump a greater
volume of blood with each beat. This is particularly true in endurance trained athletes.
Another observable response to endurance training is the lowering of resting heart-rate;
again, because of the heart’s increased stroke volume. However, changes in maximal
heart-rate are more dependent upon factors such as age and health status, rather than
training. Maximum heart-rate generally declines somewhat with age; although studies on
active athletes over many years indicate that predicted age related declines are diminished
by long-term fitness. Swimming coaches should be aware that each person may have a
slightly different maximum heart-rate during land based exercise and during swimming,
because swimming places the body in a horizontal position and the heart does not have to
work as hard against gravity. It's wise for the coach to have one maximum heart-rate as a
reference for swimming work and another as a reference for gym work.

At any submaximal swimming velocity, improvements in cardiovascular efficiency are


evident by a lower heart-rate requirement to perform that level of work. Thus, one of the
easiest ways of determining if endurance has improved over several weeks of training is
to test the swimmer on a steady-state swim (usually 10-30 minutes duration) and check
the heart-rate response. Adaptations are also shown by a decrease in the time it takes the
heart-rate to recover to a resting level following a standard workload. Faster heart-rate
recovery to a resting level indicates a positive fitness adaptation.

Respiration is the process of gas exchange that takes place in the lungs; there is also a

capacity to exchange gases (O2 and CO2) across muscle tissue. A more complete
description of the mechanisms of respiration should be taken from an exercise physiology
textbook. It's important for the swimming coach to know that the capacity for oxygen
exchange will improve as a response to aerobic training loads. Nutrition

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)

52 and health status are also major concerns because they influence one’s haemoglobin
level in the blood and iron storage capacity in blood plasma. Another of the body's
responses to endurance training is the increase in capillary density within muscle tissue.
This allows the blood to more fully profuse the tissue and thereby carries more oxygen to
ST muscle fibres, as well as removes more lactic acid from around the FT fibres. This
second effect will impact upon a swimmer’s anaerobic capacity. Endurance trained
individuals also show an increase in blood volume. The summation of all training
adaptations will affect the swimmer’s efficiency by increasing both energy supply
capacities and recovery mechanisms.

Measuring Exercise Intensity

Heart-rate has been shown to have a certain relationship with aerobic energy supply
because heart-rate must respond to a greater demand for oxygen by the muscles.
Therefore, heart-rate is often used as a marker to estimate a relative aerobic workload.
However, there are some problems inherent with both measurement error and the
theoretical significance of using heart-rate alone as a measure of intensity. When
heart-rate is taken manually by holding two fingers over a pulsating artery in the wrist or
neck; or one hand is placed on the chest to count heart beats; there can be considerable
measurement error. There is usually not enough time to count heart beats for a full
minute; therefore, a 6 or 10-second count is taken and the number of actual beats is used
to calculate a rate for one minute (i.e. by multiplying the count by either 10 or 6,
respectively). You can see that a simple miscount of one beat will mean an error of 6-10
in estimating the heart-rate. Using electronic monitoring equipment will provide a more
accurate estimate of heart-rate, but this estimate may still contain fundamental errors
because heart-rate is also sensitive to other influences. For example; body temperature,
hydration, and health status represent only three factors that may also have an influence
on heart-rate. It would be more precise to continuously monitor oxygen consumption to
estimate aerobic energy supply, but this is simply not practical for the coach. Swimming
velocity is another key marker that is generally associated with energy supply, and
velocity is easily measured with a stopwatch. Yet another marker of energy supply is the
accumulation of lactic acid in the blood. Although lactic acid is not associated with
aerobic energy supply, we must remember that both aerobic and anaerobic systems are
always working. There will be some small increases in lactic acid production as the
aerobic energy load increases, but until a critical point is reached the body’s reduction and
buffering mechanisms keep lactic acid level under control. We can see that all measures
contribute some information relevant to aerobic exercise intensity. Some of these
measures are more easily or accurately obtained than others. Because heart-rate, oxygen
consumption, and lactic acid accumulation are all physiological factors that are 'felt' or
interpreted by the individual as a perception of effort, this measure also becomes useful.
While perception of effort is not a scientific measurement, it has been shown to be a
reliable indicator of energy supply. Using a simple rating scale from 1 - 10 that represents
the effort required to swim at various speeds or intensities will give us a guide to how the
body perceives energy demand/supply. Such a scale provides the coach with a useful
estimate and can be combined with heart-rate to help define a workload.

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)
53

Scale of Perceived Exertion

Heart-rate Approx. % Lactic Acid Effort Description of bpm below Max MaxVO2
Balance Rating Perceived Effort

rest - 70 less than 30 resting level 1 very light exertion


(sub training level)

70-50 30-50 resting level 2 easy warm-up


(recovery & preparation)

50-40 50-70 removal rate 3 light to moderate work


increases (basic training effects)

40-30 70-80 both production 4-5 moderately hard & removal increase
(endurance effects) (balanced)

30-20 80-90 * both production 6-7 hard


& removal increase (endurance & speed)

20-10 90-95 * production starts 8 very hard


to exceed removal (high sub-max load)

10-max above 95 * production 9 very, very hard


exceeds removal (approaching max load)

max 100 production 10 exhaustive


exceeds removal

* Endurance trained swimmers will be able to remove lactic acid from the muscle more
efficiently and maintain a more even balance of production and removal. Therefore, lactic
acid accumulation will be slowed.

Fatigue

The term fatigue is used to describe the short-term sensation of tiredness and decline in
performance. There are a number of causes of fatigue, among them: depletion of energy
reserves, reduction of energy supply (due to the accumulation of waste products),
neuromuscular factors, and psychological factors.

As noted earlier, very short-term (for a few seconds) energy supply is dependent upon
phosphocreatine availability, and longer-term energy supply is dependent upon the
synthesis of ATP from more complex fuel sources. Lower energy requirements associated
with slow swimming speeds can be supplied from aerobic energy sources and the fuel
supply is extensive, both fats and carbohydrates can be used. Higher energy requirements
rely on glycogen as the sole fuel source for anaerobic metabolism and glycogen is also
the preferred fuel source for high aerobic energy loads.

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)

54 Therefore, glycogen depletion and replenishment in the muscle is critical. Glycogen


replenishment will be explained in the chapter on nutrition. The rate of glycogen
depletion is dependent upon the volume of intense work performed. If successive training
sessions (i.e. from one day to the next) contain work that depletes glycogen reserves, diet
alone may not be enough to bring glycogen levels back to 100%. Rest, diet and the
cycling of training intensities to draw upon other energy reserves will allow the body to
replenish glycogen in the muscle. The most common cause of accumulated fatigue at the
end of a hard training week is the incomplete restoration of glycogen. The rate of
glycogen use is an important factor to consider when the coach schedules individual
training sets. The most intense sprint sets may deplete glycogen to the point of muscle
fatigue after 15-20 minutes of interval work (note: this time estimate does not include rest
time between swims). Depending upon diet and other fatigue factors, it may take 12-24
hours (sometimes more) for glycogen stores in the muscle to return to 100% capacity.
Scheduling another hard sprint set during the ‘depleted period’ may, or may not, deliver
the specific stress the coach wants. Sometimes the coach will want the swimmer to be
‘fresh’ and ready to perform at maximum capacity and sometimes the coach will try to
‘overload’ the swimmer when he/she is at less than maximum capacity.

The accumulation of waste products from anaerobic glycolysis can produce short- term
fatigue in less than one minute of intense work. As mentioned earlier, the balance
between energy production and lactic acid removal will determine whether fatigue set in.
A well developed aerobic capacity will assist on the removal side of the balance, and
fatigue can be delayed. Muscle buffering capacity also helps to delay the onset of fatigue.
In general, we say that 'fitness' helps to reduce or delay fatigue and assists in returning the
muscles to a chemical balance where additional intense work can be performed. Repeated
intense swims as part of a training set will cause accumulated lactic acid to stay in the
muscle long after the training set has ended. This residual fatigue may cause muscle
soreness lasting after training has finished. There are techniques the swimmer can employ
to reduce this long-term fatigue, such as: massage, stretching exercises, hydrotherapy
(whirlpool bath, etc.), active and passive rest, and deep rest (i.e. sleep).

Fatigue may also result from an inability of the nervous system to activate the muscle
fibres. Electrical impulses must be transmitted across the junction between the nerve and
muscle, called the motor end-plate. Failure of nerve impulse transmission from the motor
end-plate to the muscle cell may be caused by an alteration in a number of chemical
reactions involving calcium, potassium, neurotransmitters, and/or muscle enzymes. The
stress of training or competition may change the threshold for electrical stimulation of the
muscle tissue. Fast twitch muscle fibres, because of their high threshold for stimulation,
may not receive the message to activate. Fatigue is the result of fewer FT fibres being
recruited for the required muscle action; this causes a feeling of 'weakness' in the muscle.
It's also likely that the central nervous system is another site where fatigue occurs. The
recruitment of muscle fibres depends, at least in part, upon conscious control from the
brain. Exhaustive efforts (usually from training) or anxiety (usually from competition)
may subconsciously inhibit the activation of muscles because the brain is constantly
receiving messages regarding the 'pain status' of the muscle. An increase in the perception
of effort is usually one of the early signs of fatigue. Swimmers can learn to overcome
some of these central nervous system inhibitions and thus delay this type of fatigue to
some extent.

Chapter 6 Physiological Considerations Coaching Swimming: An


Introductory Manual (second edition)
55

Psychological factors can never be completely separated from physiological factors.


Muscle fatigue and inefficient neuromuscular coordination is often brought about by
general conditions of psychological fatigue, depression, or anxiety. Keeping swimmers
happy, motivated, and focused on achieving realistic outcomes is often the best way to
delay or overcome the short term effects of fatigue.
Chapter 6 Physiological Considerations Coaching Swimming: An
Introductory Manual (second edition)
56

CHAPTER 7 TRAINING
METHODS

The previous chapter introduced basic information on how the body supplies energy to
meet the needs of swimming performance. There were also references to key training
concepts, such as: overload, recovery, specificity, variability, periodisation, and
adaptation. Now it's time to explore ways that training outcomes can be achieved through
the application of specific training methods. Then, in the following chapters, we'll
consolidate this information into logical program guidelines covering both short- term
and long-term planning.

Eight types of training, each having a specific performance outcome, are listed in the
chart and will be discussed in this chapter. The labels attached to each type of training
may vary from other textbooks, but certainly the descriptions and definitions provided
will help the coach to understand what each training method represents when encountered
under some other name. To minimise confusion, try to associate each type of training
with the energy sources, stress impact (i.e. how long will the body take to recover?),
rating of perceived exertion (RPE), and the relative proportion of race pace swimming
velocity attained. Because every swimmer's goal is to swim each race distance at the
fastest possible pace, the term 'race pace' is dependent upon the distance being swum.
Every swimmer will have many race paces, covering standard racing distances; 50 – 100
- 200 – 400 – 800 – 1500m that help set the parameters for training expectations.

Aerobic Base Training

The term, 'aerobic base', identifies a level of training that stimulates our aerobic energy
supply systems to begin to produce physiological adaptations. Therefore, aerobic base as
a training method represents swimming at a velocity that is well below any specific race
pace. It’s valuable for the heart function, respiratory and circulatory adaptations
stimulated. Because the stress of each individual swim is relatively low, to obtain an
overload affect the volume of work is usually substantial. This means that aerobic base
training loads must be sustained for a period of time with very little rest if interval
training sets are used. As the desired physiological adaptations occur our aerobic base
swimming velocity will become faster, but the relative proportion of effort will not
change. The measurement of a swimmer's aerobic base swimming velocity can be utilised
as a valuable performance indicator to reflect aerobic adaptation or basic fitness.

In theory (and indeed in practice) if a swimmer can elevate base swimming velocity,
he/she has a platform that can be used to achieve adaptations at higher levels of aerobic
energy demand. Coaches should be aware that effective training programs are ones that
work toward improvements in all areas of the energy supply spectrum. Aerobic training at
all levels along the energy supply continuum is only part of the full training package a
coach must plan. It’s a mistake to concentrate only on one method of training to the
exclusion of others, although base level aerobic training is very important (particularly at
the start of a training program).

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)
57

Training Methods

Training Energy Source(s) RPE Stress Impact Relative to Race Method (short and long-term) Pace

Aerobic aerobic metabolism 3 low - total stress is based below race speed Base on volume of work for all
distances

Aerobic aerobic metabolism is 4-7 moderate to high - total below race speed Endurance primary energy
source stress based on volume for short distances
and frequency of work close to long-
distance race pace

Critical aerobic & anaerobic con- 6-8 high - total stress usually close to race Velocity tributions are both
high, but based on frequency speed for 400m
below maximum values of application

Maximum aerobic metabolism at max, 9+ very high - total stress close to race Aerobic anaerobic
contribution based on frequency of speed for 200-
increasing use, high residual fatigue 400m events

Lactate anaerobic metabolism at 9+ very high - volume will be close to race Tolerance max, aerobic
system under self-limiting; frequency of speed for 200m
stress to help recovery use is limited by high events
residual fatigue effects

Peak anaerobic system is 8+ high to very high - total close to race Lactate stressed, aerobic system
stress based on frequency speed for 100-
increasing to meet lactate of use, residual fatigue 200m events clearance demands
may vary with volume

Sprint anaerobic (both lactic acid 6-9 moderate to very high - close to race
and alactic) systems are- total stress based on volume speed for 50m used and
frequency of use, 50m events
residual fatigue effects vary
with volume of work

Maximum anaerobic alactic system 6-8 moderate – total volume faster than 50m Speed is primarily
used (may be of work is controlled, race speed
some lactic acid fatigue is usually short- accumulation) term

Training loads in the aerobic base range will contribute to the total training volume (i.e.
the total of all types of training) at every stage of the season. In relation to higher training
loads, aerobic base loads become the 'recovery' sessions that complement more stressful
training components. The general rules for prescribing aerobic base training loads are: (1)
maintain submaximal heart-rate around 50 bpm below maximum, (2) swim with
minimum rest as continuous work or very short-rest intervals, (3) sustain the workload
over a sufficient period of time, usually in excess of 20 minutes and sometimes up to 60+
minutes, (4) maintain efficient swimming technique, and (5) even pace or slightly
negative split each swim. **

* Negative split swimming means the second half of the interval is swum faster than the
first half.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)

58 Coaches should use aerobic base training as an additional opportunity to improve


stroke technique and racing skills, such as streamlining off the wall, turns, and finishes.
Done correctly, aerobic base swimming is not always easy, because it involves great
effort to maintain the type of stroke technique that will be used at race speeds. Correct
technique includes maintaining the timing, kicking tempo, and body position desired
during peak performance.

Coaches normally monitor a full range of performance improvements over a season.


There should be parallel changes in aerobic base, aerobic endurance, sustained speed, and
peak speed. If aerobic base improves without improvements in other race components,
then the coach is probably concentrating too much on low intensity training and more
work should be done at higher training speeds. If aerobic base training speed decreases
(i.e. becomes slower), it's most often the result of detraining or failing adaptation caused
by illness, injury, or too much intense work. Each case represents a lack of balance in
program design.

Examples of Aerobic Base Training Sets

Set #1 (mixed interval distances)


2 x [150m / 300m / 450m / 300m /150m] first cycle free & second cycle formstroke
or form/free mix, rest 10 seconds between swims (check HR) total 2700m

Set #2 (also mixed distances)


2 x [4 x 50m / 3 x 100m / 2 x 200m / 300m] as above, but rest 5 seconds after 50's,
total 2400m (a swimmer using the same pace as in the set above will take longer to
compete this set, because of more frequent breaks -- keep the rest short)

Set #3 (fixed interval distances)


10 x 250m or 10 x 300m freestyle, rest 10 seconds between swims, swimming at
1:30 pace (for example), this set will take about 40-47 minutes (note the use of ‘off’
interval distances, such as 250m, adding variety to the set) total 2500-3000m

Set #4 (over-distance intervals, freestyle)


3 or 4 x 1000m freestyle, rest about 15-20 seconds (this could also be done as an
arms only pulling set using a pull-buoy) total 3000-4000m

Set #5 (over-distance, formstroke)


5 or 6 x 600 backstroke of breaststroke, rest about 15-20 seconds (also try this set
for breaststrokers using fins and a ‘wave-action’ body motion following each pull,
or alternate 600 breaststroke swim and 600 breaststroke with fins) total 3000-3600m

Set #6 (single long swim)


3000m straight freestyle swim (a variation could be 15 x 200 IM consecutive
swims, no rest between change-overs) if this type of set is used be careful that
swimmers maintain the correct pacing and insist upon correct technique (which
includes turns) total 3000m
Chapter 7 Training Methods Coaching Swimming: An Introductory
Manual (second edition)
59

Aerobic Endurance Training

As the term suggests, this type of training is designed to stress the swimmer so that
greater volumes of work can be performed. It builds upon the physiological adaptations
stimulated by aerobic base training, but endurance work is performed at a greater
swimming velocity. Building endurance is important not only to the distance swimmer,
but to formstroke specialists and sprinters. Increasing one's aerobic endurance by
delivering more oxygen to the working muscles has the effect of enhancing the
swimmer’s ability to recover from stressful anaerobic efforts. Adaptation to endurance
training allows the swimmer to swim further/faster while relying on a major proportion of
energy supply from aerobic sources.

Aerobic endurance outcomes are similar to aerobic base training, except that greater
depletion of muscle glycogen occurs and neuromuscular fatigue is greater. Endurance
training is performed with heart-rates in a range of 50-30 beats below maximum. Aerobic
endurance training relies on three components to stimulate successful adaptation: (1)
sufficient volume to overload the aerobic energy system, (2) controlled intensity, usually
within a limited heart-rate range, and (3) relatively short rest between interval swims. As
with aerobic base training, work of 20 minutes or more is required. This is usually made
up of interval type sets of swims, although continuous long swims may be used. Interval
training methods provide us with an almost unlimited degree of variation that can be
applied to constructing training sets. The distance of each interval swim may remain the
same or vary up/down, strokes may change, specialised kicking only or pulling only sets
will also produce endurance outcomes provided the three components are present.
Working through a range of heart-rates also allows the coach to specify descending times
and other variations in the training set.

The amount of rest programmed between repeat swims should be carefully planned to
overload the aerobic energy system. The amount of time spent swimming during an
interval will probably be six to eight times the amount of time spent resting between
swims. For example, swimmer 'A' performs 30 x 100m holding on average time of
1min.28sec. (i.e. 88 seconds) with a heart-rate around 35 bpm (± 5) below maximum.
Using a 6:1 ratio means about 15 seconds rest between swims. It may be convenient for
the coach to round-off the interval time to 1min.40sec. (this actually gives the swimmer
only 12 seconds rest on average), but the standard interval gives the coach greater control
if there are several swimmers sharing the training lane. Swimmer 'B' performs 10 x 400m
maintaining an average swim time of 6min; a ratio of 8:1 means that each swim will
commence on a 6min.45sec. clock cycle. In general, as the interval distances become
greater the coach should select a greater work-rest ratio. Almost any repeat distance can
be incorporated into endurance sets; however, there may be a tendency for swimmers to
exceed the heart-rate range if very short intervals (i.e. such as 50m) are used. There is
some research to indicate that the best endurance adaptation occurs when rest is
controlled between 10 and 30 seconds.

Anaerobic threshold is reached somewhere in the range of heart-rates used for aerobic
endurance sets. There are various tests used to estimate AT pace, usually they consist of a
single distance (1500-3000m) performed at the ‘fastest possible even pace’. The average
100m velocity on the test swim provides an estimate of the velocity where anaerobic
threshold is reached. Chapter 7 Training Methods
Coaching Swimming: An Introductory Manual (second edition)

60 References to target heart-rates used during the various types of training are worth
further explanation at this point. While it's generally true that most of the athletes we train
will have similar maximum and resting heart-rates, the actual individual values from
swimmer to swimmer may vary considerably. Maximum heart-rate is influenced by
factors such as age, sex (i.e. girls generally have maximums 5-10 beats/minute higher
than boys), and genetic profile. While resting heart-rate is influenced by adaptation to
aerobic base training, health status, and general state of fatigue. The difference between
maximum and resting heart-rates is the range in which we prescribe exercise intensity
(remember that heart-rate is only one indicator of intensity). Clearly, two swimmers each
performing at a heart-rate of 160 beats/minute (bpm) may be under different stress.
Swimmer 'A' might be working at 20 bpm below maximum, while swimmer 'B' might be
45 bpm below! It's also likely that each swimmer is training at a different number of
beats/minute above resting values. Because maximum heart-rate is a relatively stable
measure, while resting heart-rate is sensitive to change, we prescribe training heart-rate
based upon beats below maximum. Coaches often ask "if bpm below max is the training
goal, how do I start by determining a swimmer's maximum heart-rate". Simple,
administer a test such as the one shown on the following page and record each swimmer’s
results.

Test to Determine Individual Maximum Heart-Rate


There is no single 'best' test to use. However, these criteria form the basis for a test the
coach may construct:
• heart-rate will increase rapidly at the start of an exercise effort of high intensity, but
don’t expect max heart-rate after the first swim,
• it may take 2-3 minutes for some individuals to reach their maximum heart-rate,
• heart-rate will decline rapidly (particularly in well conditioned swimmers) if the
intensity of the effort is reduced or broken (i.e. when the swimmer slows or is given too
much rest between swims),
• achieving maximum swimming speed over a short distance is not always directly
related to attaining maximum heart-rate.

Suggested Test Protocol After completing a warm-up routine - swim 200m increasing the
pace on each 50m from 'easy' to 'moderately-hard' - check heart-rate during a 10-15
second break after the 200m effort - swim 2 x 100m high effort repeats (as close to race
pace as possible) with very short rest between each (3-5 seconds is enough) - check
heart-rate immediately upon completion of the second 100m effort. 1. if the swimmer's
second 100m effort is significantly slower (i.e. more than 5-6 seconds) than the first
100m effort, then repeat the test substituting 3 x 50m efforts (3 seconds rest between) for
the second 100m effort, measure heart-rate after the final 50m effort. 2. an electronic
heart-rate monitor should be used (i.e. for accuracy) / young age- group swimmers may
not be experienced enough to accurately use palpation methods to measure heart-rate;
therefore, the coach must develop his/her skill in taking heart-rates manually.

Although the ideal situation is to keep the heart-rate in a target range, in practice the
coach and swimmer will notice a gradual upward drift in heart-rate during the middle

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)
61

and later stages of any aerobic set. Cardiac drift may occur early in the training set if the
swimmer has a poor endurance base or the intensity is too high during the first few
swims. It’s normal for swim times to remain constant and yet heart-rate will gradually
climb, in this situation the coach has two courses of action. First, the coach can encourage
the swimmer to maintain swim times while the heart-rate continues to drift upward; this
results in a greater anaerobic energy contribution. Second, the coach can advise the
swimmer to slow slightly and keep the heart-rate under control; this keeps the energy
supply primarily within the aerobic range. A third possible option, change the send-off
interval to increase rest between swims, is a less desirable choice if endurance is the
primary objective of the training set. By allowing more rest between repeat swims,
recovery is enhanced and endurance overload is sacrificed. However, if maintaining
swimming pace is important, then changing the interval to allow slightly more rest may
be an acceptable option. In this case a greater percentage of the required energy will come
from anaerobic sources.

Examples of Aerobic Endurance Training Sets

Set #1 (fixed interval distance)


15 x 200m or 30 x 100m freestyle, rest 15 seconds between swims, or use a send-off
interval to allow this rest (variation - descend times on swims 1-3, 4-6, 7-9, 10-12,
13-15, etc.) total 3000m

Set #2 (also fixed distance)


60 x 50m mixed strokes, rest 5-10 seconds (or appropriate send-off interval), swim
15 repeats each stroke or alternate 10 x 50 formstroke, 5 x 50 freestyle, check
heart-rate every 10th swim, total 3000m

Set #3 (mixed interval distances)


4 - 8 x (100 - 200 - 300m) freestyle, rest 10/20/30 seconds (or establish an interval
time per 100m such as 1min.25sec., i.e. 1.25/2.50/4.15 send-off times), total
2400-4800m

Set #4 (mixed distances & descending times)


5 - 10 x (50 - 100 - 150m) formstroke, rest 5/10/15 seconds (or establish interval
time per 50m), on the second set swim 50m faster than first set, on third set swim
100m faster than first set, on fourth set swim 150m faster than first set, on fifth set
swim all three distances faster than the first set (repeat sequence if extending the
work to 6-7-8 etc. sets), total 1500-3000m

Critical Velocity Training

This type of training is based upon a heart-rate – swimming velocity response that
predicts maximum heart-rate is reached at the same point where maximum oxygen
consumption is reached. In practice swimmers can only maintain maximum oxygen
consumption for a few minutes duration (naturally this capability is extended with
specific training). In theory if we can train at a level just below the point where max HR

and max VO2 are reached, a specific training load can be sustained long enough to create
an overload effect.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)

62 The relationship between swimming velocity and heart-rate is roughly linear at sub-
maximal rates. When maximum heart-rate is reached, there is a logical limitation on
oxygen delivery to the working muscles because the heart can not pump blood any faster.
Maximum heart-rate is a physiological parameter not sensitive to changes with training,
but the heart's stroke volume (i.e. the amount of blood pumped with each beat) is subject
to change. Therefore, endurance training that approaches maximum heart-rate will
overload the heart to stimulate increases in stroke volume. Critical velocity training has
also been called 'heart-rate sets' by some authors, although any type of aerobic training
can include some evaluation of heart-rate as one measure of intensity. The key to
successful training at high, yet sub-maximal, loads is to identify a 'critical velocity' for
repeat swims when the heart-rate is still 15-10 bpm below max (some coaches use a range
of 20-10 bpm below max).

Very little has been said about lactate production thus far in the discussion of training that
is primarily aerobic. As pointed out in the discussion on energy systems, lactic acid is
produced at all levels of exercise intensity. Energy is always being produced via both
aerobic and anaerobic energy pathways, the relative proportions of each are determined
by the energy demand and one’s training adaptations. At very high aerobic loads (such as
critical speed training), lactate levels will be high but remain stable (i.e. not continuing to
climb). If lactic acid levels in the muscle are stable during critical speed training sets, it
means the mechanisms for lactic acid removal are being stimulated to keep pace with the
increased production. This is an important feature about critical velocity training; it forces
the body to change the dynamics of lactate removal as well as providing a high level of
aerobic stress.

As well as the benefits derived from critical velocity training, there are some limitations.
Because of the higher overall stress placed on the body, this type of training can not be
done every day. There are also concerns that young swimmers could be over-exposed to
this type of training and fail to recover sufficiently for adaptation to take place. Other
concerns are the volume of critical velocity work per individual training sessions and the
accumulated volume during a week or fortnight’s training. The ideal amount of rest taken
between repeat swims is somewhat variable, any two swimmers may need slightly
different amounts of rest. These general guidelines will help the coach plan an
appropriate amount of critical velocity training:
• volume of CV work applied at any one time should be in a range of 1500-3000m
(somewhat below this range during a taper period),
• volume of CV work should build up during the season and then reduce during the taper,
• CV training is generally applied no more than twice per week for senior age-group
swimmers,
• interval distances of 100 - 200m are best, although intervals ranging from 50m to 300m
may be used in some cases,
• rest intervals of 35-50 seconds are usually prescribed; this may be extended to one
minute in some cases,
• CV work may be performed in freestyle or formstrokes; the training outcomes become
more specific if the swimmer's racing stroke is used,
• CV work should be introduced with target heart-rate of 20 bpm below max; if the
swimmer is able to hold swim velocity at the target heart-rate then training loads
approaching 10 bpm below maximum can be used.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)
63

It will be necessary for the coach to monitor the progress of swimmers during the CV
training set to maintain the target heart-rate. It's also an advantage for the coach to
predetermine a swimmer's critical speed and review this about every 4-6 weeks to
determine a new critical velocity. A useful formula has been developed to allow the coach
to estimate critical velocity based on two test swims.

Estimating Critical Velocity

Test Protocol: Critical velocity can be estimated from two criterion values. Freestylers
will use times for 50m and 400m freestyle and formstrokers will use times for 50m and
200m. It's not advisable to use competition times for this purpose. Prior to the test the
swimmer should be reasonably fresh (i.e. not showing symptoms of residual fatigue from
previous training sessions) and the coach should prepare the swimmer with an adequate
warm-up. From a 'push start' swim 50m at maximum effort, record the time. Following
this effort a sufficient recovery swim (or set of swims) should be performed. Again from
a 'push start' the swimmer performs a maximum effort 400m freestyle (or 200m
formstroke) swim, record the time.

Calculation: CV = (400m - 50m) ÷ (400 time* - 50 time*)


* all times are given in seconds (i.e. 5min.30sec = 330sec)

Example: 400m time 5min.42.61sec. 50m time 31.27sec.


CV = (400-50) ÷ (342.61 - 31.27) = 1.124 m/sec

Training Estimate: 100m repeat swims @ 1.124 m/sec


(i.e. 100m ÷ 1.124 m/sec = 88.9sec or 1min.28sec)

Set Construction: 20 x 100m on 2min. interval, HR 10-20 bpm below max.


target time of 1min.28sec allows about 32sec rest

Variations: Critical Speed sets may involve 150m or 200m repeat swims.
Calculate the target velocity and add a fraction of a second for
150m repeats and 1-2 seconds for 200m repeats.

Coaches will find that CV becomes faster after a few weeks due to adaptation. During CV
training sets if the heart-rate begins to climb to maximum there are two options. First, ask
the swimmer to slow slightly to maintain HR at the target level. Second, increase the time
interval within the guidelines for set design (i.e. provide rest that is closer to 50sec. to one
minute). If these modifications are made and the heart- rate still continues to approach
maximum, then the training set should be shortened or stopped. Swimmers having poor
endurance capacity will find it difficult to cope with CV training. If training heart-rate
remains below the target zone the coach should not shorten the rest, but encourage the
swimmer to go faster. This may be an indication that the targeted critical velocity needs to
be revised. Over time, the coach will become familiar with the way individual swimmers
respond to this type of training and be able to adjust training parameters on an individual
case basis.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)

64 Maximum Aerobic Training

Training at one's maximum aerobic capacity can only be maintained for a few minutes
duration. When aerobic energy supply is at its limit and heart-rate is at maximum, any
additional demand for energy must be supplied from anaerobic (lactic acid producing)
energy sources. This causes a rapid imbalance in the production / reduction of lactic acid
in the muscle and creates a rapid increase in lactic acid accumulation. This type of
training is very stressful and, depending upon the volume of work done, may require two
or three days before full recovery is complete. When we talk about 'recovery' in this
context we don't mean full rest, but a reduction in specific training stress to allow residual
fatigue to dissipate.

There is some debate regarding the effectiveness of this type of training for pre- pubertal
swimmers. Prior to and during the childhood growth spurt it may not be advisable or
necessary to program maximum aerobic training. Most of the gains in maximum aerobic
capacity during childhood are the result of growth and maturation. Sub-maximal aerobic
training loads are more than sufficient to stimulate continued improvements among young
swimmers.

However, maximum aerobic training is a useful training method for senior age-group and
elite level swimmers. As with other forms of highly stressful training a swimmer's ability
to absorb the training stimulus and adapt to it will depend to a large extent upon other
fitness components. Certainly there is a readiness for this type of training that’s
determined by how well a swimmer recovers and adapts to other components of the
training program. The volume of each training load, and frequency of application, will

determine the amount of overload applied during a Max VO2 set. General guidelines for
prescribing maximum aerobic training are:
• total volume of work (per session) will be much lower than other types of training,
800-1500m is a realistic target,
• highly stressful performance should be maintained for 3-5 minutes, then some level of
recovery will be required,
• recovery should always take the form of active recovery (i.e. swimming at aerobic base
level).

Examples of Maximum Aerobic Training Sets


Set #1 (freestyle)
3 x (4 x 100m max effort, rest 5-10 sec. between * / swim 200m recovery following
the fourth 100m effort) heart-rate should reach maximum during the second 100m
effort and remain at that level on the third and fourth swims, total 1200m effort and
600m recovery

Set # 2 (formstroke)
3 x (3 x 100m max effort, rest 5-10 sec. between * / swim 150m recovery following
the third 100m effort) heart-rate should reach maximum and remain there during the
short breaks (recovery to ‘true’ resting level is never achieved during the recovery
swim), total 900m effort and 450m recovery

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)
65

Set #3 (freestyle of IM)


4 x (400m free or IM max effort, swim 150-200m easy between each effort) time
allocated to recovery swim should equal half the time of the effort swim, total
1600m effort and 600-800m recovery

* Fixed time intervals may be used if the swimmer can hold all repeat times within
the interval.

Because we are demanding maximum aerobic contribution, the corresponding swimming


velocity will be fairly high and blood lactate levels will be increasing. It's very difficult to
maintain the required high swimming velocity for 3-5 minutes of sustained effort with
heart-rate at maximum. Therefore, broken swims with very short rest between each
segment are used. Each 3-5 minute period of effort will be followed by an active recovery
lasting approximately 50-60% of the effort time (i.e. two and one-half to three minutes
recovery time). Well conditioned distance and middle distance swimmers will be able to
achieve maximum oxygen consumption quickly. The time frame usually allows
300-400m of effort, followed by 150-200m recovery swimming. Training overload is
provided by repeating this cycle a number of times.
Lactate Tolerance Training

This type of training is done with the intention of producing very high amounts of lactic
acid in the muscle. The swimmer experiences the feeling of muscle soreness, burning,
and fatigue and responds by adapting to the physical and psychological demands. When a
swimmer attempts to swim at/close to peak velocity from the start of a swim, the energy
demand is very high from the start. The body relies heavily on anaerobic metabolism to
supply the required high level of energy. Therefore, the process of lactic acid
production/reduction is out of balance very quickly. Lactic acid is normally removed from
the muscle tissue via the blood, but the fast twitch muscle fibres being called upon do not
have a direct blood supply. Therefore, lactic acid accumulates in the muscle and results in
a characteristic feeling of 'muscle burn'. Lactic acid will diffuse across the FT muscle cell
membranes and eventually be carried away from the tissue. The body responds to the
condition of lactic acid accumulation by increasing the buffering capacity in the muscle
and blood to neutralise more lactic acid. Changes also occur in the production of some
enzymes. In addition, there may be a psychological adaptation to the characteristic
discomfort felt in the muscles; the swimmer can learn (to some extent) to tolerate this
pain and continue slightly longer at a very high level of effort.

As with maximum aerobic training methods, there is good rationale for not including
lactate tolerance training as a primary training method used among pre-pubescent
swimmers. Although the notion that children do not produce high lactates has been shown
to be incorrect, the qualitative response (in terms of muscle chemistry) among children to
this type of training may not produce the desired outcome. Young swimmers will
experience the feeling of high levels of lactate in the muscle during peak lactate training
and if properly motivated will achieve the desired psychological outcome.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)

66 By its very nature, lactate tolerance training is very stressful and produces residual
fatigue which may last for 48-72 hours (i.e. 2-3 days). Although most of the lactic acid
will be removed from the muscle within 30-60 minutes, higher than resting levels may
remain for days. Several factors will influence the rate at which the muscle chemistry
returns to normal. First, the swimmer's aerobic fitness will influence the rate of recovery
because of greater lactic acid removal capabilities. Second, recovery enhancement
techniques such as massage, spa baths, sauna, etc. will help to speed the recovery process.
The fast twitch muscle fibres used during high velocity swimming may also sustain
micro-trauma (i.e. tearing or damage) during the state of acidosis induced during this type
of training. This may take several days to repair so coaches should be aware that this is a
natural part of the stress-recovery-adaptation cycle. However, when there is repeated high
stress training loads, the swimmer may not be able to cope and adaptation fails to occur.
It's for this reason that lactate tolerance training should not be used too frequently. These
general guidelines will help the coach use lactate tolerance training appropriately:
• total volume of work (per session) will be limited, between 800-1200m is a realistic
target,
• individual swims must be long enough to produce a peak blood lactate level,
• a series of peak lactate performances are scheduled so that successively higher levels of
residual blood lactate are present at the start of each swim,
• active recovery swimming between efforts should be 3-5 times longer than the effort
swim,
• the adaptation period must be long enough for the swimmer to absorb the training
before it is repeated; the length of this period will depend on the amount/types of other
training in the total program.

Muscle lactate will presumably reach a saturation level, but blood lactate will continue to
increase throughout the set. The net effect is that a high level of lactic acid must be
tolerated in the muscle tissue. Since the training objective is to achieve peak lactate
production on each swim, repeat distances of 75-200m are recommended for lactate
tolerance sets. Sprint specialists, because they have a high percentage of fast twitch
muscle fibre, may be able to effectively use shorter repeat swims on lactate tolerance sets.

Examples of Lactate Tolerance Training Sets

Set #1
8 x 100m max effort on 4-6min interval, swim easy between efforts, total 800m
(excluding recovery swimming)

Set #2
5 x 200m max effort on 8-13 min interval, swim easy between efforts, total 1000m

Set #3 (sprinters)
3 x (100m max effort on 5-6min + 4 x 50m max effort on 2-3 min), swim easy
between efforts, total 900m
Note: freestyle or formstroke may be performed on any of these sets

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)
67

Peak Lactate Training

This type of training helps the swimmer to adapt to race conditions over 100-200m. It
may also be called race simulation or quality training because the physiological response
should be similar to that encountered in competition. A training overload is created when
more than one swim is performed. This type of training also gives the swimmer an ideal
opportunity to practice the desired stroke technique at the desired stroke rate and count,
race start, turns, and pacing. Guidelines for the construction of peak lactate training sets
are relatively simple:
• each swim should be approached with race intensity; post-race heat-rate should be at
maximum,
• total volume of work (per session) should be limited to 400-600m (distance specialists
may need to increase this total)
• active recovery swimming should be encouraged between effort swims (note: this is a
good opportunity to practice post-race swim-down routine),
• there should be a suitable adaptation period between sessions containing this type of
training,
• competition events may be used in the same way that a peak lactate set is used (i.e.
several race efforts over a 2-3 hour period with swim-down between each).

Examples of Peak Lactate Training Sets

Set #1
4 - 6 x 100m race effort on 20+min. interval, recovery swimming or aerobic base
training set between efforts, total 400-600m of race quality (excluding recovery
swimming)

Set #2
3 x 200m race effort on 30min. interval (as above), total 600m of race quality
Set #3
3 x 200m race effort (each effort split, 5-10sec. rest at 100m and 150m) on 30min.
interval (total swim time for 100-50-50 faster than race time), total 600m of race
quality

Note: freestyle or formstroke may be performed on any of these sets

Sprint Training

Specialist training for 50m sprint events should stimulate both alactic and lactic acid
producing anaerobic energy production. Coaches should note that pre-pubertal age
swimmers will utilise a higher percentage of aerobic energy even on sprint swims. The
key elements of racing capacities that are developed by using sprint training are energy
supply and neuromuscular patterning to specific race requirements (i.e. the rate of muscle
contraction). Coaches should always encourage the best possible technique during this
type of work.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)

68 As with other types of training, the overall fitness of swimmers will influence their
ability to absorb larger amounts of sprint training. Sprint training sets should include
these elements:
• partial recovery between effort swims to stimulate an overload (i.e. there will be a
gradual build-up of lactic acid in the muscle),
• because this is primarily an anaerobic exercise load, heart-rates after each swim are
expected to be high (although they are not necessarily at maximum),
• work to rest ratios range from 1:1 to 1:3 or slightly more,
• active recovery swimming may be used between efforts,
• individual repeat distances of 25-75m (usually the distance of each repeat swim will be
below the range of peak lactate accumulation from a single swim),
• training volume (per session) is dependent upon maintaining swimming speed,
800-2400m of work is realistic.
Variety can be added to training sets by programming descending times, ascending or
descending distances, and slight changes of pace to try to negative split swims. Rest
between swims may be a combination of active recovery swims and complete rest.

Examples of Sprint Training Sets

Set #1
24 x 50m on 1min.15sec. to 1min. 30sec interval (or sufficient rest within an
appropriate work-rest ratio), total 1200m of fast swimming

Set #2
64 x 25m on 45sec interval, sprint odd numbered swims / build-up even numbered
swims, approximately 1000m fast swimming and 600m at lower intensity

Set #3
10 x (25 - 50 - 75m) on 45sec/1min.30sec./2min. interval (rest or easy 25m swim
between efforts), total 1500m of fast swimming

Note: freestyle or formstroke may be performed (send-off interval adjusted to fit the
stroke and ability of the swimmer)

Sprint training is perhaps less stressful in the short-term than previously mentioned
maximum loading types of work because partial recovery is provided between swims.
However, because this type of training relies primarily on anaerobic metabolism there
will be accumulation of lactic acid in the muscle that may result in residual muscle
soreness. The body will also draw heavily upon glycogen stores in the muscle as a fuel
source, so glycogen levels will be depleted after this type of training. Depending upon
conditions such as diet, immediate glycogen replenishment (i.e. use of a sports drink
during the training session), and other types of training programmed during the session;
glycogen stores may not return to full capacity for several hours. This may impact on how
frequently this type of training is scheduled.

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)
69
Maximum Speed Training

Short bursts of energy are required at times during all race distances. Explosive
movements are used during the push-off following each turn, as well as the increased
stroke rate while swimming in-out of turns. In sprint events (i.e. 50m) the requirement for
maximum speed from the first stroke is obvious. Maximum velocity swimming over a
short distance will call upon alactic anaerobic energy supply. Repeated maximum, but
short duration efforts, stimulate adaptation from both energy supply mechanisms and the
neuromuscular system. Actual swimming speed may be greater than the average speed
over 50m race performance. This type of training will rely on the recruitment of fast
twitch muscle fibres. All swimmers, regardless of their preference as a distance or sprint
specialist, need this type of training regularly. Sprint swimmers will benefit most in terms
of the contribution made to specific improvements in race performance. Guidelines for
prescribing maximum speed sets, also called ‘High Velocity Overload’ (HVO) are:
• maximum swimming velocity, stroke rate, stroke length, and correct technique are the
most important factors (heart-rates are irrelevant),
• individual repeat distance is usually 25m or less,
• work to rest ratios should favour rest; 1:4 or more is appropriate,
• active rest (i.e. low intensity swimming) may be used between explosive efforts,
• total volume of high-speed swimming (per session) should be relatively low, 400- 600m
is appropriate,
• small amounts of maximum velocity work may also be combined with drill work, kick
or pull sets (i.e. as part of the active recovery).

Benefit from maximum speed training sets is achieved when swimmers are reasonably
fresh (i.e. not exhausted from other types of training) and focused on quality
performance. However, there is evidence to suggest that performance benefits are also
achieved from adaptations in neuromuscular control (as opposed to adaptations of the
energy system) when maximum speed sets are scheduled at the end of a training session.
Residual fatigue effects are not unusually high, provided that sufficient recovery
swimming is used during and after the required efforts.

Examples of Maximum Speed Training Sets (HVO Sets)

Set #1
16 x 25m on 1min.10sec. interval (or sufficient rest within an appropriate work- rest
ratio), active recovery swimming between efforts, total 400m (exclusive of recovery
swimming)

Set #2
9 x (3 x 50m ) → first 50m sprint 20m and 30m easy swim / second 50m build- up
20m and sprint 15m, then 15m easy / third 50m build-up 30m and sprint 20m
1min.20sec. interval, approximately 500m of fast swimming

Set #3
8 x (4 x 25m) on 1min. interval → after each cycle swim 200m recovery, total 800m
(exclusive of recovery swimming)

Note: freestyle or formstroke

Chapter 7 Training Methods Coaching Swimming: An Introductory


Manual (second edition)

70 Each type of training can be defined in terms of the physiological stress that is applied.
However, the net affect of training is a complex mixture of volume, intensity, and
specificity (i.e. type of training applied). Similar training programs will not automatically
produce the exact same training result for any two swimmers. Well constructed training
programs contain the right mixture of training sets that are designed to: (1) achieve the
desired training objectives over both short and long-term timeframe, (2) complement each
other for the best possible adaptation, and (3) provide an overload without subjecting the
swimmer to overtraining.
Chapter 7 Training Methods Coaching Swimming: An Introductory
Manual (second edition)
71
CHAPTER 8 DRILLS and
STROKE-RATE

Stroke drills attempt to isolate some aspect of stroke technique and then pattern a more
desirable movement. Coaches should always remember that swimming performance is
determined by skill level as much as fitness level. If the skills emphasized in a drill are to
become successfully integrated into a swimmer’s stroke, they must be practiced regularly
and practiced under increasingly stressful conditions. Therefore, drills are not meant to be
easy, they may be just as demanding (but in a different way) as other components of the
training program.

When selecting a drill and then applying it in a training program, the coach should
consider these points:
• Is the drill appropriate to the current skill level of the swimmer?
• What are the complexities of the movements?
• Will the physical demands of the drill require low or high energy expenditure, physical
strength, or specific range of motion?
• Does the drill have application to one or more strokes?

There are three phases of progression in learning a drill. The first is to learn the
movement pattern at a slow tempo. Neuromuscular patterns and kinaesthetic awareness
may take some time to develop into a routine or automatic action. During the second
phase the movement speed is increased so the drill is performed at a rate similar to
swimming speeds. During the third phase, movement speed is combined with pressure;
that is, some demand is placed on the swimmer to perform the drill within a timeframe.
Pressure may also be applied by performing the drill under an external load; such as
wearing a drag suit, hand paddles, fins, or towing an object. It’s important for the
swimmer to maintain the correct movement pattern while under pressure.

There are an infinite number of drills that a coach could use. However, most coaches
maintain a limited number of drills they feel are useful. In some ways, teaching a new
drill is similar to learning a new stroke, it will take time for the swimmer to perfect the
drill and move from phase-one into phases-two/three.

The Green Licence Coaching Course placed considerable emphasis on stoke progressions
and the use of drills to shape desirable stroke patterns. Rather than repeat a list of drills in
this text, we will try to provide additional background information on the rationale for
drill selection and use. The coach will use his/her experience and other resources
(manuals, video, and demonstration) to build a personal set of drills.

When teaching drills coaches need to emphasize the key elements of control, precision,
rhythm, and co-ordination along with the targeted movement pattern. To achieve the
desired results, drills must be performed under close supervision with adequate feedback.
Ideally the swimmer-to-coach ratio should be lower when performing drills than in most
training situations. Those coaches who would dismiss the use of drills from a training
program might argue that drills represent ‘ideal’ movements and have little connection
with the complex set of movements used in

Chapter 8 Drills and Stroke-Rate Coaching Swimming: An


Introductory Manual (second edition)

72 swimming. As with all aspects of training, the model comes from the ideal and the
application (i.e. commitment on the part of the swimmer) will determine the results
achieved. Therefore, the coach must have a rationale behind each drill so that it ends up
having a purpose and specific outcome. In this way drills can be used as a prescriptive
tool to fine tune a stroke pattern.

The term drill is applied to countless variations of teaching techniques and coaching
practices having a wide range of performance outcomes. This author would like to
propose that the term drill is further broken down into four categories. Each category
represents a somewhat different performance outcome.

1. Stroke Development Drills

Movement patterns and techniques designed to develop the stroke from the basics
through to the whole stroke, performed under competitive conditions. The major focus of
stroke development drills is to enhance the positive performance aspects of technique.
These drills are ideally practiced just prior to a 'main training set' to review correct
movement patterns and maximise the transfer of the stroke model into a performance
setting.

This concept of stroke development is generally expressed by a series of movement


progressions that build in their complexity, increase in their intensity, and maximise
propulsive efficiency. Stroke progressions that have been developed as part of Australia’s
National Junior/Youth Program serve as a good example of stroke development drills.
During each drill sequence the swimmer must concentrate on distance per stroke and
streamlining the body.

Variations on these progressions involve sensory perception and efficient application of


propulsive force. For example, drills that incorporate hand paddles or fins are used to
change the level of sensory input. They stimulate neuromuscular feedback so that stroke
patterns are reinforced. However, these drills by themselves do not necessarily correct
fundamental stroke faults when the whole stroke is swum. Stroke correction may require
a more specific application of drills.

2. Stroke Correction Drills

These drills use techniques designed to correct specific stroke faults (i.e. significant
deviations from the stroke model). They work to correct a negative aspect of stroke that is
causing the swimmer to swim with decreased efficiency. These drills should not focus on
cosmetic changes (i.e. technique that only makes the swimmer look better), but they
should be applied to fundamental defects that limit the swimmer's performance potential.
For example, a slight spread of the fingers is less important to the propulsive affect of a
stroke than a major defect such as dropping the elbow during the middle of a stroke. The
essential coaching question to ask is: "Will the stroke correction drill increase speed by
improving efficiency?" If the answer is ‘no’, leave it alone and concentrate your coaching
efforts elsewhere.

Chapter 8 Drills and Stroke-Rate Coaching Swimming: An


Introductory Manual (second edition)

You might also like