You are on page 1of 69

DEGREE PROJECT IN VEHICLE ENGINEERING,

SECOND CYCLE, 30 CREDITS


STOCKHOLM, SWEDEN 2021

Development and construction of a


mechanically sprung shock
absorber with adjustable spring
stiffness for mountain bikes

MARTIN HOLM

KTH ROYAL INSTITUTE OF TECHNOLOGY


SCHOOL OF INDUSTRIAL ENGINEERING AND MANAGEMENT
TRITA 584

www.kth.se
Examensarbete TRITA-ITM-EX 2021:584

Utveckling och konstruktion av en mekaniskt


fjädrad stötdämpare med justerbar styvhet för
terrängcykling

Martin Holm
Godkänt Examinator Handledare
- Ulf Olofsson Ulf Olofsson
Uppdragsgivare Kontaktperson
Intend BC Cornelius Kapfinger

Sammanfattning
I detta examensarbete utvärderas möjligheten att använda en mekanisk fjäder för att uppnå
justerbar fjäderstyvhet hos en stötdämpare avsedd för terrängcykling. En mekanisk fjäder ökar
stötdämparens känslighet jämfört med en fullt justerbar luftfjäder. I dagsläget finns på marknaden
ingen mekanisk fjäder vilken kan erbjuda tillräcklig justeringsmån för att passa cyklister med stor
viktvariation. Därför kan en mekanisk fjäder med ett brett styvhetsspann vara banbrytande om en
sådan kan tillämpas på dagens terrängcyklar.
Arbetet har utförts som ett produkt- och konceptutvecklingsprojekt där den slutliga
konceptdesignen har optimerats analytiskt och verifierats numeriskt. Den fjäder som tagits fram
har integrerats i en ny dämparkonstruktion och stötdämparsystemet har dimensionerats för att
passa dagens terrängcyklar.
Resultatet är en prototyp av en stötdämpare med en fjäder som passar cyklister mellan 70-88kg.
En alternativ fjäder passande cyklister mellan 59-75kg har också tagits fram. Eftersom en justerbar
fjäder vars design möjliggör användning i en stötdämpare för dagens terrängcyklar har slutsatsen
dragits att en justerbar mekanisk fjäder kan fungera inom detta tillämpningsområde. Det är inte
möjligt med dagens material att utforma en mekanisk fjäder med tillräckligt justerbar styvhet för
att passa alla åkare. Det är däremot möjligt att använda en fjäder som passar cyklister inom ett
viktspann på omkring 15-17kg. Detta är mellan 70–110% mer än vad liknande produkter
tillgängliga på marknaden idag kan erbjuda.

Nyckelord: Justerbar styvhet, fjäderdämpare, terrängcykling

1
2
Master of Science Thesis TRITA-ITM-EX 2021:584

Development and construction of a mechanically sprung


shock absorber with adjustable spring stiffness for
mountain bikes

Martin Holm
Approved Examiner Supervisor
- Ulf Olofsson Ulf Olofsson
Commissioner Contact person
Intend BC Cornelius Kapfinger

Abstract
In this thesis project, the possibility to make use of a mechanical spring to achieve stiffness
adjustability in a shock absorber for mountain bikes is evaluated. A mechanical spring increases
the shock absorber’s sensitivity compared to a fully adjustable air spring. Today, there are no
mechanical springs available on the market that offer enough stiffness adjustment to suit different
riders with large variation in weight. Therefore, a mechanical spring with a wide stiffness
adjustment range could be ground-breaking if it is possible to implement in mountain bike shock
absorbers.
The work has been carried out in accordance with a product- and concept development approach
where the final concept design has been optimised analytically and verified numerically. The
developed spring has been integrated in a new damper design and the complete damper and spring
system has been dimensioned to fit current mountain bike frames.
The result is a prototype shock absorber with a spring to suit riders between 70-88kg. An
alternative spring for cyclists between 59-75kg has also been proposed. Since these springs have
been made to fit current mountain bikes, it was possible to conclude that a mechanical spring with
a wide range of adjustable stiffness is feasible for mountain bike application. With available spring
steels, it is not possible to accommodate every rider with only one spring. It is however possible
to achieve adjustment that is suitable for a rider weight range of roughly 15-16kg. This is between
70-110% more than similar products available on the market can offer.

Keywords: Adjustable stiffness, coil shock, mountain bike

3
4
FOREWORD
Here I thank the persons and companies that have been helpful and/or offered special support for
completing this Master thesis.

I would first like to thank Cornelius Kapfinger, CEO of Intend BC, for offering me the opportunity
to take on this project. It has been a goal of mine since the first day of my university degree to one
day make an MTB shock absorber. In completing this project, I have achieved this goal, and by
offering me the chance to do it, Cornelius has truly made my dream come true. He has also offered
continuous support and taught me skills that I will find useful for the rest of my career. This despite
running Intend BC on his own and undoubtably having plenty of work to keep him more than
busy. So, Cornelius – Thank you!
There are of course other honourable mentions I must make. Ulf Olofsson has provided great
guidance as my academic supervisor throughout the project. Special thanks must also be given to
Ralph Geraets from Con-Sept and Dave Fielding from Force Technology for offering very helpful
advice regarding spring design and being very patient in discussions about manufacturing.
Last, but not least, I owe many thanks to my family and friends who have continuously supported
me through my degree and allowed me to come to this point.

Martin Holm

Älvsjö, October 19

5
NOMENCLATURE
Lists of notations and abbreviations used in this Master thesis are presented in this chapter.

Notations
Symbol Description
E Young´s modulus
𝐺 Shear modulus
r Radius
t Thickness
p Pressure
V Volume
mc Mass of cyclist
mf Mass of front end of bike
Nf Normal force front wheel
Nr Normal force rear wheel
𝐹𝑠𝑝 Spring force
g Acceleration due to gravity on earth
𝑘 Spring stiffness
C Spring Index
Kw Whaal Correction factor
𝐻0 Free height of spring
𝐻𝑏 Bind height of spring
𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 Stack height of inactive sections of spring
Δ𝐻 Spring stroke
𝑁𝑎 Number of active spring coils
𝐷𝑡 Spring wire diameter
𝐷𝑚 Spring mean diameter
𝐷𝑖 Spring inner diameter
𝜏 Shear stress
𝛾 Helical spring coil pitch
𝛼 Archimedean spiral pitch
𝑅𝑚 Tensile strength

6
Abbreviations
CAD Computer Aided Design
MTB Mountain Bike
LR Leverage Ratio
RWT Rear Wheel Travel
CR Compression Ratio
IFP Internal Floating Piston
LSC Low Speed Compression
LSR Low Speed Rebound
FE Finite Element
FEA Finite Element Analysis
FEM Finite Element Method
UTS Ultimate Tensile Strength

7
LIST OF FIGURES

Figure 1. Full suspension MTB (Santa Cruz Megatower). (Image sourced from Santa Cruz bikes [1] and labels added
by the author) ________________________________________________________________________________ 14
Figure 2. Coil shocks from two large global manufacturers. a) RockShox Super deluxe coil [2], b) Fox DHX [3]. ___ 15
Figure 3. Commercially available air shocks. a) RockShox Super Deluxe Ultimate, b) Fox FloatX _______________ 16
Figure 4. Cutaway of a common air shock with labelled elements. (The image is sourced from amb magazine [5],
and labels added by the author of this paper) _______________________________________________________ 17
Figure 5. Cutaway of a monotube coil shock with labelled elements. (The image is sourced from Blister [6], and
labels added by the author of this paper). __________________________________________________________ 17
Figure 6. The 'Jack Spring’™ concept proposed by Hollander et. Al [12]. __________________________________ 18
Figure 7. The Sprindex adjustable-rate coil soring with adjustment operation [14]. _________________________ 19
Figure 8. Diagram of example bike with positions of mass centres and important dimension for calculating spring
stiffness. The of the bike is taken from Santa Cruz Bicycles [1] __________________________________________ 21
Figure 9. Free body diagram of example bike. Only forces necessary for spring stiffness calculations have been
included. _____________________________________________________________________________________ 21
Figure 10. System simplification process for finding spring force based on rider weight, bike geometry and average
linkage leverage ratio. __________________________________________________________________________ 22
Figure 11. Assembly drawing of the 2021 Fox Float X2 shock with general dimensions ______________________ 25
Figure 12. Sketch of adjustable straight coil concept with an upper support matching the coil pitch. ___________ 27
Figure 13. Search algorithm for finding constrained optimum in adjustable coil spring design. ________________ 29
Figure 14. Estimated available space for spring and spring retainers. ____________________________________ 30
Figure 15. Representation of coil spring with variable diameter at uncompressed state ______________________ 32
Figure 16. Coil spring with variable diameter at bind height ____________________________________________ 32
Figure 17. Sketch of torsion spring concept implemented in an MTB shock. _______________________________ 35
Figure 18. Idealised spring force curves for dual chamber air springs with the equal pressures and different chamber
size ratios. ___________________________________________________________________________________ 36
Figure 19. Dual chamber air spring curves for a 40% chamber volume ratio, same pressure, and different
compression ratios. ____________________________________________________________________________ 37
Figure 20. Concept sketch of air/coil hybrid spring ___________________________________________________ 38
Figure 21. Spring curves for different configurations of the air/coil hybrid spring concept. ____________________ 39
Figure 22. Section views of an early version of the recirculating damper design with components marked out. ___ 45
Figure 23. Available space for adjustable coil spring assembly. _________________________________________ 46
Figure 24. Valve piston alternatives and their port sizes._______________________________________________ 47
Figure 25. Flow through additional channels and valves in the recirculating damper. ________________________ 48
Figure 26. a) Chamber pressures in the damper during compression and rebound with small pistons and 100psi IFP
pressure. b) Damping force over shaft speeds with small piston and 100psi IFP pressure. ____________________ 48
Figure 27. a) Chamber pressures in the damper during compression and rebound with small pistons and 100psi IFP
pressure. b) Damping force over shaft speeds with small piston and 100psi IFP pressure. ____________________ 49
Figure 28. von Mises stresses in upper spring retainer from finite element simulation on final version.__________ 50
Figure 29. Back view of simulation results with maximum stress area visible ______________________________ 50
Figure 30. Minimum general dimension for upper spring retainer to give enough safety for yield ______________ 51
Figure 31. Final shock and spring assembly with spring adjusted to a) minimum stiffness=70N/mm and b) maximum
stiffness=87N/mm. ____________________________________________________________________________ 53
Figure 32. Section view of final design. _____________________________________________________________ 53
Figure 33. Section view of final damper valving and flow channels. ______________________________________ 54
Figure 34. Assembly drawing of final shock and spring assembly with general dimensions and shock stroke. _____ 54
Figure 35. Constraints and loads in verification model ________________________________________________ 56
Figure 36. Results from convergence study with spring at minimum stiffness setting and maximum stroke with a
mesh size of a) 10mm, b) 5mm, c) 2mm, d) 1mm. For all results, a curvature refinement ratio of 0.5 was used. __ 57
Figure 37. Verification simulation results for a) minimum stiffness and 32.5mm stroke, b) minimum stiffness and
65mm stroke, c) maximum stiffness and 32.5mm stroke, d) maximum stiffness and 65mm stroke. ____________ 58

8
LIST OF TABLES

Table 1. Weight percentiles of the Swedish population 2010-2011 [15] ___________________________________ 20


Table 2. Weight percentiles for the Swedish population 2018-2019. Source: SCB, "Undersökningarna av
levnadsförhållanden (ULF/SILC)" [16] ______________________________________________________________ 20
Table 3. Frame data from sample models for finding desired spring stiffness adjustment range. ______________ 24
Table 4. Optimised spring parameters and estimated maximum stresses for two different spring steels. ________ 31
Table 5. Optimised spring parameters and estimated maximum stresses in a coil spring with variable wire thickness
for two different spring steels. ___________________________________________________________________ 34
Table 6. Results from qualitative concept evaluation of the four initial adjustable spring concepts. ____________ 40
Table 7. Explanation of evaluation categories and outline of scoring systems ______________________________ 42
Table 8. Targets, limits, directions of improvement, weight and importance values for each category used in the
concept desirability scoring ______________________________________________________________________ 43
Table 9. Numerical category scores for each concept _________________________________________________ 44
Table 10. Weighted desirability scores and overall system scores for each concept. _________________________ 44
Table 11. Final optimised spring parameters for different stiffness ranges using OTEVA 70 SC spring steel. ______ 52
Table 12. Results from mesh convergence study _____________________________________________________ 56
Table 13. Comparison of FEM stress to design model stress.____________________________________________ 58
Table 14. List of calculated stresses for comparison between standard coil spring and adjustable prototype spring 59
Table 15. Results from spring design optimisation. ___________________________________________________ 60

9
TABLE OF CONTENTS

FOREWORD 5

NOMENCLATURE 6

LIST OF FIGURES 8

LIST OF TABLES 9

TABLE OF CONTENTS 10

1 INTRODUCTION 12

1.1 Background 12
1.2 Purpose 12
1.3 Delimitations 13
1.4 Method 13

2 FRAME OF REFERENCE 14

2.1 Mountain bikes 14


2.2 MTB shock absorbers 15
Shock architecture 15
Shock size standards 16
2.3 Shock Architecture and damping 17
2.4 Adjustable mechanical springs 18

3 CONCEPT DEVELOPMENT 20

3.1 Pre-study 20
Population weight data 20
Stiffness adjustment range 20
General shock dimensions 24
Usability 25
3.2 Requirement Specification 25
Product goal 26
Requirements 26
3.3 Concept generation 26
Straight coil spring 26
Coil spring with variable wire diameter 31
Torsion Spring 34
Coil/air hybrid spring 35

10
3.4 Concept evaluation 40
Qualitative evaluation 40
Quantitative Evaluation 41
Choice of Concept 45
3.5 Detailed design 45
Shock and damper construction 45
Damper modelling 46
Spring adjustment feature 49
Optimisation of final spring design 51
Final CAD assembly 53
3.6 Test strategy 55
Model verification 55
Establishing effects of spring stress in MTB application 55
3.7 Analysis and model Verification 55
Numerical verification using FEA 56
Comparison with RS coil spring 59

4 RESULTS 60

4.1 Spring stiffness adjustment 60


Most suitable spring type for an adjustable stiffness MTB shock 60
Achievable adjustment range 60

5 DISCUSSION AND CONCLUSIONS 62

5.1 Discussion 62
The spring 62
Shock integration 63
5.2 Conclusions 64

6 RECOMMENDATIONS AND FUTURE WORK 65

6.1 Recommendations 65
6.2 Future work 65

7 REFERENCES 66

11
1 INTRODUCTION
This chapter aims to describe the background to the project along with a problem description and
a proposed research question. The chapter is concluded by presenting the purpose of the study,
delimitations in the project and finally outlining the method followed to answer the research
question.

1.1 Background
MTB suspension has been drastically improved over the last three decades. Today, several
companies offer shock absorbers and suspension forks suitable for all types of riders and riding
styles. Most such products consist of two fundamental components – a spring and a damper acting
in parallel.
The function of the damper is to dissipate energy from impacts and to slow down and control
movement. This is usually achieved by hydraulic circuits and orifice or shim stack dampers. On
the other side, there is the spring. This component stores impact energy and controls the static
riding position.
In almost all modern damped performance shock absorbers, the spring is either pneumatic or a
mechanical coil spring. Pneumatic springs offer great adjustability and low weight. However, the
pneumatic spring usually requires at least one more dynamic seal which increases friction. It is
also difficult to tune the spring to achieve a zero-force offset to initialize shock movement.
Therefore, riders looking for the most sensitivity often opt for coil sprung shocks. Coil springs
without preload do not introduce any friction from the spring side and initialise movement at
exceptionally low loads. Their disadvantage is greater weight and lack of adjustability as the
stiffness of a given coil spring cannot be changed. There are some products that allow stiffness
adjustment in a limited range, but a problem remains that several springs with different stiffness
must be produced for any coil shock to accommodate all possible riders. Also, riders cannot easily
change the behaviour of their coil shock without changing the entire spring.
Intend BC therefore wishes to explore the possibility of developing a fully adjustable coil shock
which suits a wide range of, if not all, riders. That is the task for this thesis project and the research
question that must be answered is:
Is it possible to design a mechanical spring with adjustable stiffness that is suitable for use on
mountain bikes by riders with a wide weight range?

1.2 Purpose
Modelling and real-world prototype testing is required to definitively tell if a fully adjustable
mechanical spring is a feasible product for mountain bikes. The resulting knowledge from this
project should show whether significant improvements can be made to the current mechanically
sprung shocks. If it is not possible, then fully adjustable mechanical springs are likely not feasible
for mountain bikes. Thus, this thesis will be the basis for a decision from Intend BC whether to
start production of an adjustable mechanically sprung shock.

12
1.3 Delimitations
• In depth damper tuning is not a part of the thesis project. The damper unit of the shock is
designed based on already existing parts and rough damper modelling was made to choose
the best piston option.
• The spring is only designed for a rear shock with a standard installation length of 230mm
and a stroke from 57.5-65mm.
• The target rider weight ranges are based on the adult Swedish population and the desired
stiffness adjustment ranges are aimed to accommodate for “a majority” of the population.
• Modelling of springs performed aims to find the theoretical optimal spring design that
corresponds to the most desirable stiffness range. The design choice for prototyping
however is largely based on the stiffness requirements of potential test riders.
• Detailed, non-linear modelling of the shock in the rear linkage of bikes is not part of the
development. A simplified linkage model with assumption of a static rider position is used.
The use of this model together with data from some common bike models is aimed to give
an estimate of desired stiffness adjustability ranges for the established target rider weight
ranges.
• Spring fatigue life estimates have been left out of this project due to lack of load data for
MTB shocks and excessive uncertainty in pure estimates of such.
• Due to prototype manufacturing delays and the time limit for the thesis work, physical
testing of the shock and spring have not been included in the project as initially planned.
A suggested test procedure is included instead.
• The spring element is intended for use on mountain bikes, the choice of concept for
prototyping is therefore based not only on performance but also manufacturability, price,
space constraints, usage environment and usability.

1.4 Method
A product development approach was used to answer the research question. As such, the work was
divided into six stages:
1. Pre-study
2. Requirement specification
3. Concept generation
4. Concept evaluation and choice
5. Detailed modelling, construction, and manufacturing
6. Prototype testing, model verification, and product validation
The information gathered during the pre-study served as the basis for a preliminary requirement
specification drafted with input and wishes from the industrial supervisor. Also, information and
data from the pre-study laid a foundation for the concept generation stage.
A limit of four concepts was set for the concept generation stage due to the time constraint of the
thesis.
Analytical spring models and some numerical simulation data were used to evaluate the concepts.
Also, input from the industrial supervisor regarding manufacturability and usability was strongly
weighted in the evaluation process. Weighted desirability scoring was used to quantitatively
compare the concepts and choose the most promising for prototyping.
After choosing a concept in conjunction with the industrial supervisor, detailed modelling of the
spring was performed using Matlab and FEA in Solid Edge 2020 [1]. In parallel, a damper was

13
constructed in which the spring would be mounted and used. The construction was carried out
using Siemens’ CAD software Solid Edge 2020.
During the detailed modelling stage, the spring parameters were optimised to minimise spring
stress given space constraints set by the damper and available bike frame space.

2 FRAME OF REFERENCE
This chapter presents a summary of knowledge relevant for the development of an adjustable
spring for mountain bikes. Brief summaries of mountain bikes, current shock absorbers and
existing concepts for adjustable mechanical springs are included.

2.1 Mountain bikes


Mountain bikes (MTB) are pedal powered bikes adapted for off-road use and designed to handle
rough terrain. The differentiation compared to road bikes mainly lies in the use of wheel
suspension on mountain bikes. Hardtails only feature front wheel suspension while full suspension
bikes have both front- and rear wheel suspension. The rear wheel is attached to the main frame by
a linkage that allows relative movement. The linkage drives a shock absorber (shock).

Figure 1. Full suspension MTB (Santa Cruz Megatower). (Image sourced from Santa Cruz bikes [1] and
labels added by the author)
Figure 1 shows a full suspension MTB with labelled main components. Different categories of
mountain bikes exist, and they are broadly categorised by the amount of wheel travel and intended
use. Enduro bikes are made for very rough and steep terrain and generally feature between 150-
180mm rear wheel travel. The bike in figure 1 belongs to this category and these bikes are

14
optimised for descending while maintaining some ability to pedal upwards. Suspension
performance is therefore often prioritised slightly higher than weight in this MTB category,
although weight is still and important factor.

2.2 MTB shock absorbers


Shock architecture
MTB shocks are very similar in architecture and design as shock absorbers found on motorcycles
and heavier motorised vehicles. For mountain bike application, they are generally smaller as they
must fit in the bike frame and generally experience lower loads.

A shock consists of two elements: a spring and a damper. These are acting in parallel and the
spring stores energy while the damper dissipates energy. For MTB, the current types of springs
used are either mechanical helical coil springs or air springs.

Figure 2. Coil shocks from two large global manufacturers. a) RockShox Super deluxe coil [2], b) Fox
DHX [3].
Two coil shocks from two large manufacturers are shown in figure 2. The spring is positioned
concentric to the damper shaft and held in place with an upper and a lower spring retainer. This
configuration offers no adjustment of the spring stiffness for individual tuning. The stiffness is
instead adjusted by exchanging springs that usually are produced in increments in stiffness of 4-
5N/mm. Due to the extremely low break-away force of the coil spring, coil shocks offer high
sensitivity, but the lack of stiffness adjustment makes them hard to fine tune for individual
preference.

15
In contrast, two air shocks from the same manufacturers are shown in figure 3.

Figure 3. Commercially available air shocks. a) RockShox Super Deluxe Ultimate, b) Fox FloatX
The air spring is composed of an airtight canister that is concentric to the damper shaft. A piston
inside the canister compresses the air causing an increase in pressure and thus an increase in spring
force. As the air pressure of the spring is easily adjusted through an external valve, it is very easy
for a user to precisely tune the stiffness to their individual preference. The drawback is an
additional dynamic seal in the air spring compared to coil shocks. This increases friction and
reduces shock sensitivity. Since the seal friction depends on seal squeeze which in turn depends
on the sealing pressure, the sensitivity of the air shock is reduced with increased spring stiffness.

Shock size standards


Some standard sizes exist in the MTB industry which all frame manufacturers work around. The
latest and current standards are the so called “Metric” and “Trunnion” sizes. In the Metric sizes,
the shock is mounted in the frame through a top and a bottom eyelet which are free to rotate on
bushings. The trunnion shocks feature a lower eyelet, but the upper part of the shock is attached
to the frame by bolts that are allowed to rotate on ball bearings in the frame. In this way, all frames
are designed to take the axial force introduced when the shock is compressed, but not bending or
torsional moments.

For enduro bikes, two metric and two trunnion sizes are the most common. These are [4]:

• 230 x 57,5-65 – Metric size with an eyelet distance of 230mm and a stroke of 57,5-65mm
depending on frame specification.
• 210 x 47,5-55 – Metric size.
• 205 x 57,5-65 – Trunnion size with a build-in distance of 205mm and a stroke of 57,5-
65mm.
• 185 x 47,5-55 – Trunnion size.

16
2.3 Shock Architecture and damping
The common coil- and air shocks feature similar principal damper architecture. The major
difference is that the spring piston in the air spring is usually connected to the damper shaft and
the spring is an integral part of the shock. Figures 4 and 5 show cut views of an air- and a coil
shock respectively.

Figure 4. Cutaway of a common air shock with labelled elements. (The image is sourced from amb
magazine [5], and labels added by the author of this paper)

Figure 5. Cutaway of a monotube coil shock with labelled elements. (The image is sourced from Blister
[6], and labels added by the author of this paper).
When the damper shaft is compressed, so is the spring which exerts an opposing force on the
spring plate or air piston. The damper piston also pressurizes the fluid in the damper compression
chamber. This generates an oil flow through the damper compression valves to the damper oil
reservoir. To compensate for the volume by the damper shaft inserted in the shock body, the
volume of the oil reservoir is increased by an Internal Floating Piston that is displaced. The IFP is
backed by an air-filled pressure chamber that pressurises the damper and allows oil to flow back
through the damper valves as the damper shaft retracts during the rebound stroke.

17
During the compression stroke, oil must also flow to the backside of the damper piston, to the
rebound chamber. In a monotube damper, this is allowed by ports and valves in the damper piston.
In a twin tube or recirculating damper, the damper piston can be solid, and oil is allowed to flow
through a separate path from the reservoir to the backside of the piston.

Damper tuning is commonly achieved by varying shim stack configurations in the damper valves.
The valves restrict the flow but allow variable orifice restriction depending on damper shaft speeds.
Thus, the shim stack stiffness can be altered to tune the damper curve. Modelling of the damper is
generally a complicated task due to the involved geometries and number of interdependent
variables. Some models have been proposed in the literature which have been shown to be more
or less accurate [7] [8] [9] [10]. Generally, damper tuning is done by feel and experience, but
during damper design, modelling is important to avoid cavitation and spikes dues to over-damping
and to achieve desirable damper adjustment.

2.4 Adjustable mechanical springs


In robotics, where the need for precise linear actuation is required to handle variable loads,
adjustable stiffness mechanical springs are more common than in the vehicle sector. For robotics,
research is ongoing to achieve such stiffness adjustment in various ways. In a thesis paper, Kiliç
gives an overview of springs with adjustable stiffness for use in robotic and suggests a new design
achieving adjustable stiffness of a non-linear torsion spring using cams [11]. Most of the spring
systems reviewed rely on opposing springs actuated by levers or cams. By adding pretension to
antagonistically working springs connected by a roller, the torsional stiffness of the roller can be
altered. The systems allow fine tuning, but the mechanisms require space that makes it extremely
difficult for them to be adapted for MTB shocks.

Another way to achieve adjustability is proposed by Hollander et. Al [12] where the active section
of a coil spring can be adjusted. This is achieved by adjusting the position along the spring of one
of the spring supports and a concept sketch is shown in figure 6.

Figure 6. The 'Jack Spring’™ concept proposed by Hollander et. Al [12].


Their developed concept is named adjustable ‘Jack Spring’™ and is aimed for use as an adjustable
mechanical tendon in a robotic actuator system. However, there is promising potential for the same
principle to be applied in MTB shocks thanks to the limited number of required components. It is
not certain it can be directly applied though since an MTB shock would likely require the use of a
larger portion of the spring’s available stroke subjecting the spring to higher stress. This is
something that is not investigated in Hollander’s paper.

18
Similar ways to achieve adjustment of various configurations of coil springs have also been
proposed for, and sometimes implemented on, road vehicles. Doulatani et. Al gives a good
overview of such technologies [13]. These include using stoppers to alter the stroke of series
connected springs, coil springs with various thickness tapered plates to limit the compression
between coils, prestressing series connected springs of different stiffness, compressible bushings
in series with a coil spring, threaded spring retainer to alter the number of active coils and a
combination of series and parallel springs activated by a pivot mechanism. Since most of these are
aimed for road vehicle application, the systems have greater space requirements than what is
possible for MTB shocks. Recurring though is the adjustment of the number of active coils in a
spring which does not require much space.

One adjustable spring is commercially available for mountain bikes today. This is the Sprindex
coil spring which works on the principle on changing the number of active coils [14]. An image
of the Sprindex is shown in figure 7.

Figure 7. The Sprindex adjustable-rate coil soring with adjustment operation [14].

This spring is designed to fit most coil shocks on the market and is sold as an after-market product
only. It offers an adjustment range of 40-50lb/in or approximately 7-8,75N/mm and is aimed for
riders who roughly know their required spring rate but want the option of fine tuning their spring
rate. Because the Sprindex must fit a majority of already existing shocks, it could be that a wider
adjustment range could be achieved with a more integrated shock and spring design.

For a 65mm stroke Sprindex with an adjustment range of roughly 79-88 N/mm, the specified
weight is 446 g. Depending on the rate, springs for the same shock stroke range from 390-486 g
[14].

19
3 CONCEPT DEVELOPMENT
In this chapter the working process of developing an adjustable spring concept is described. The
chapter outlines the process following a chronological order starting with the initial pre-study and
ending with the analysis of the final concept design.

3.1 Pre-study
Population weight data
For defining a target weight range for the adjustable spring developed in this project, the Swedish
population was used as a basis. The reason for focusing on Sweden alone is partly due to the author
being Swedish, but also due to simplicity finding population data with the Swedish Central Bureau
for Statistics, SCB. Weight data presented from SCB regarding the adult population of Sweden
(16 years and older) was given in quartiles for the entire population as well as for men and women
in each age bracket. The most current data available was from 2010-2011. After contacting the
statistics bureau, some data from 2018-2019 was also provided although it had not yet been
published. The population weight percentiles are presented in the tables below.

Table 1. Weight percentiles of the Swedish population 2010-2011 [15]


Weight in kg
Percentiles 90% 75% 50% 25% 10%
Median
In total 95 85 74 64 57
Men 100 90 82 74 67
Women 84 75 65 59 54
Age:
16-19 years 83 74 65 58 52
20-29 years 92 81 71 62 55
30-39 years 96 85 75 64 57
40-49 years 97 86 77 66 58
50-59 years 97 88 78 68 60
60-69 years 95 86 76 67 60
70-79 years 92 83 75 65 58
80+ years 85 78 70 60 54

Table 2. Weight percentiles for the Swedish population 2018-2019. Source: SCB, "Undersökningarna av
levnadsförhållanden (ULF/SILC)" [16]
Wight in
kg
Percentiles 10% error 25% error 50% error 75% error 90% error
Median
All 16-64 56,5 0,4 64 0,4 75 0,3 85,5 0,4 98 0,9
years
Men 65,5 0,7 73,5 0,5 82 0,5 92 0,7 102,5 1,1
16-64
years
women 16- 53 0,5 59 0,4 65 0,3 75 0,6 85 1,1
64 years

Stiffness adjustment range


Suitable spring stiffness for an MTB shock is dependent on rider weight, frame kinematics, frame
geometry, body positioning and riding style. The two latter are difficult to generalize as they vary
for all riders and depend on track types, weather conditions and other uncontrollable factors. The
dynamic position and weight distribution also changes constantly during riding. A desired stiffness

20
adjustment range for the spring being developed in this project was therefore based on a static rider
position.

During descents, a rider is predominantly in a balanced standing position on the bike. The heels
are kept down, the hips are slightly shifted towards the rear wheel, no contact is made with the
saddle and the body weight is mainly placed on the pedals. A slight portion of the body weight is
put on the handlebars for steering. This puts the rider weight centred over the bottom bracket.
Therefore, it was assumed in this project that the mass centre of a rider is placed on a vertical line
above the bottom bracket the bike.

Figure 8. Diagram of example bike with positions of mass centres and important dimension for
calculating spring stiffness. The of the bike is taken from Santa Cruz Bicycles [1]

Figure 9. Free body diagram of example bike. Only forces necessary for spring stiffness calculations have
been included.

21
The compressive force on the rear shock depends on the sprung mass and the weight distribution
between the front and rear wheel. The weight distribution is found by solving the equilibrium
equations from the free body diagram of a simplified rider and bike system as presented in figure
9:

𝑁𝑟 + 𝑁𝑓 − (𝑚𝑐 + 𝑚𝑓 )𝑔 = 0, (1)

𝑚𝑓 𝑔𝑊𝐵
𝑁𝑓 𝑊𝐵 − 𝑚𝑐 𝑔𝐶𝑆 − = 0. (2)
2

It was assumed that the weight of the rider, 𝑚𝑐 𝑔, acts only through the bottom bracket as displayed
in figure 8 and 9. The sprung weight of the bike itself was also included based on rough estimates
of general bike weight and position of mass centre. It was assumed that the sprung mass of the
bike itself was positioned roughly in the middle between the front and the rear axle. This was
denoted, 𝑚𝑓 , for front centre mass and it was estimated as the mass of the entire bike minus the
wheels, tires, suspension fork, rear derailleur, and cassette.

For calculating the spring force in the rear shock, the normal force on the rear axle, 𝑁𝑟 , is of
importance. This is found by solving equations 1 and 2:

𝑊𝐵−𝐶𝑆 𝑚
𝑁𝑟 = 𝑚𝑐 𝑔 ( + 2𝑚𝑓 ), (3)
𝑊𝐵 𝑐

where 𝑚𝑐 is the mass of the cyclist. Here, the weight of the unsprang mass, i.e., wheels, tires, fork,
cassette, and rear derailleur are not included since it does not contribute to the compression of the
shock during static loading.

Depending on the linkage design, full suspension mountain bikes have different leverage curves
throughout the rear wheel travel. The leverage ratio between the rear wheel and the rear shock is
rarely constant but often decreases towards the end of the wheel travel. This makes it very time
consuming to find the exact spring stiffness for a given sag on many different bikes. To find a
suitable stiffness range for a selection of representative bikes; their average leverage ratio was
instead used. This is calculated by dividing the Rear Wheel Travel, 𝑅𝑊𝑇, as stated by the
manufacturer, by the shock stroke, 𝑠𝑡𝑟, specific for the frame:
𝑅𝑊𝑇
𝐿𝑅 = . (4)
𝑠𝑡𝑟

Figure 10. System simplification process for finding spring force based on rider weight, bike geometry
and average linkage leverage ratio.

22
The statically loaded bike was simplified by modelling the rear linkage as a lever bar with the
calculated rear wheel load and a linear spring acting on each side of a single pivot. Since the
average leverage ratio was used, it was also assumed that the leverage ratio was constant. With
this simplified model, presented in figure 10, together with equation 3 and 4, the spring force was
derived:
𝑊𝐵−𝐶𝑆 𝑚
𝐹𝑠𝑝 = 𝐿𝑅 ∗ 𝑚𝑐 𝑔 ( 𝑊𝐵 + 2𝑚𝑓 ). (5)
𝑐

More accurate linkage modelling would have given more accurate estimates of the spring force
for any given bike. However, the development of an adjustable spring should be suitable for as
many different bike frames as possible and therefore only an estimate of the spring force was
necessary. The simplified model approach was also motivated by the fact that the spring being
developed should be adjustable. Regardless of the detail in the modelling step, spring stiffness is
ultimately a personal preference, and the theoretical ideal setting often deviates from the real
customer need. Thus, for this thesis it was beneficial to establish only an estimated required
adjustment range for a several bike frames for a given rider weight bracket. Since the spring is
adjustable, this would allow the user to come closer to the ideal setting, regardless of bike without
over complicating the modelling process.

With the spring force, the required stiffness given a rider weight and desired sag could then be
derived from Hooke’s Law:
𝐹𝑠𝑝
𝑘 = 𝑠𝑎𝑔∗𝑠𝑡𝑟. (6)

The sag is the desired compression of the suspension when the bike is statically loaded on a flat
surface and is usually stated as a percentage. This is the typical measurement used for finding the
correct spring stiffness for mountain bikes and usual the range is 25-35%.

Seven different bike models from established brands were chosen to represent the vast offering of
modern enduro bikes available on the market. The criteria for the bikes were:

• Rear wheel travel between 150-180mm,


• Built for a rear shock of the size 230-57,5-65mm,
• Not delivered with a dual crown fork (to avoid downhill bikes),
• 27.5” or 29” wheels.

From each model, relevant geometric data was taken for size M and L since these sizes fit most
riders. For each model and size, the average leverage ratio was calculated, and the collected data
is presented in table 3.

23
Table 3. Frame data from sample models for finding desired spring stiffness adjustment range.
Bike Travel Shock stroke Average Lev. Chainstay Wheelbase
[mm] [mm] Ratio [mm] [mm]
Megatower M 160 57,5 2,78 435 1207
Megatower L 160 57,5 2,78 435 1231
Nomad M 170 62,5 2,72 430 1222
Nomad L 170 62,5 2,72 435 1256
Clash M 170 65 2,62 434 1238
Clash L 170 65 2,62 434 1265
Meta AM29 M 160 62,5 2,56 433 1258
Meta AM29 L 160 62,5 2,56 433 1285
Capra 29 M 180 65 2,77 435 1203
Capra 29 L 180 65 2,77 435 1227
Slash M 160 62,5 2,56 437 1222
Slash L 160 62,5 2,56 437 1264
Remedy M 150 57,5 2,61 435 1167
Remedy L 150 57,5 2,61 435 1206

The data was collected from the manufacturer websites (Megatower and Nomad from Santa Cruz
Bicycles [1], [17], Clash and Meta AM29 from Commencal bikes [18], [19], Capra 29 from YT-
industries [20], Slash and Remedy from Trek Bicycle Corporation [21], [22]). It was assumed that
the total weight of every bike was 15kg and that the sprung weight was roughly 7.5kg. The desired
sag range was chosen as 28-33% to allow for some margin of error.

Based on the population weight data, desired target weight brackets were defined. These were
ranked best/ideal to acceptable from their range:

• Ideal: 54-100kg to cover the middle 80% of both men and women.
• Very good: 59-90kg to cover the middle 50% of both men and women.
• Good: 64-85kg to cover the middle 50% of the overall population.
• Acceptable:
o 74-90kg to cover middle 50% of men.
o 59-75kg to cover middle 50% of women.

For each weight bracket, the required spring stiffness adjustment range was determined by finding
the maximum allowable spring stiffness for the lightest rider and the minimum allowable stiffness
for the heaviest rider. Thus, the minimum spring stiffness would give no less than 28% sag for the
lightest rider on the bike with the lowest leverage ratio while the maximum stiffness would give
the heaviest rider no more than 33% sag on the bike with the highest leverage ratio. Using eq. 5
and 6 along with the representative bike data and the target weight brackets, the following target
adjustment ranges were established:

• Ideal: 54-98 N/mm (54-100kg)


• Very good: 59-89 N/mm (59-90kg)
• Good: 64-84 N/mm (64-85kg)
• Acceptable:
o 73-89 N/mm (74-90kg)
o 59-75 N/mm (59-75kg)

General shock dimensions


Available design space for the shock was based on the Fox Float X2 air shock. This is one of the
larger shocks currently available on the market. It often come as the original shock on many

24
complete bicycles and fits most frames. It is also a shock which dimensions can be directly
retrieved from the manufacturer which makes them reliable. An assembly drawing is shown in
figure 11.

Figure 11. Assembly drawing of the 2021 Fox Float X2 shock with general dimensions

From the dimensions of the Float X2 shock, the following maximum shock dimensions could be
established:
• Maximum shock body diameter: 57,7mm.
• Maximum shock width: 97,5mm.
Usability
Ideally, spring adjustment should be done with as few adjuster operations as possible. Current air
shocks typically rely on a single valve to adjust air pressure and the Sprindex is adjusted through
the turn of one dial. Thus, it was identified that achieving stiffness adjustment through a single
dial or adjuster would be ideal.

3.2 Requirement Specification


Some overall goals regarding performance and usability were set as the basis for spring concept
and shock development. These were then condensed into verifiable requirements in a preliminary
requirement specification. During the entire thesis project, this requirement specification was then
referenced to evaluate and verify the different concepts and the final prototype.

25
Product goal
The overall goal of the product was to allow full adjustability of spring stiffness without the need
to exchange springs or add preload. This was to be achieved without sacrificing sensitivity as this
would be the benefit of a mechanical spring over already adjustable pneumatic springs. The
adjustment range was to be suitable for a wide variety of riders and enduro type mountain bikes.

Requirements
Functional:
• Adjustable to give a sag of 28-32% for a rider weight range of:
o Ideally:
o Preferably:
o At least:
o For bikes in the Enduro category.
• Rebound and compression damping must be externally adjustable to accommodate the
entire stiffness adjustment range.
• A mechanical spring element must be included.
Geometrical:
• The shock must have an installation length of 230mm.
o The stroke must be adjustable between 57,5-65mm.
• The external diameter of the spring cannot be greater than 57,7mm.
• The widest dimension of the shock cannot exceed 97,1mm.
Usability:
• Stiffness adjustment should be possible without removing the shock from the bike with
exception for unusual frame designs.
• Stiffness adjustment should not be achieved with more than one dial.

3.3 Concept generation


Five concepts were developed based on different types of mechanical springs and a mechanical
pneumatic hybrid. Ways to achieve stiffness adjustability was first considered before space
constraints and suitability for an MTB shock were looked at. Before in depth modelling of the
concepts was started, an initial qualitative assessment of the concepts was conducted. If damper
system integration, usability or suitability for MTB environment were not considered easily
achievable, the concept was not taken further. Analytical stiffness and stress models were used to
estimate spring performance and required dimensions after the qualitative assessment.

Straight coil spring


For a straight coiled compression spring, the spring stiffness is given by:

𝑑4 𝐺
𝑘 = 8𝐷3𝑡 𝑁 , (7)
𝑚 𝑎

where 𝐷𝑡 is the wire diameter, 𝐺 is the shear modulus, 𝐷𝑚 is the mean coil diameter and 𝑁𝑎 is the
number of active spring coils [23]. Once fitted in a shock, the wire diameter, mean diameter or
shear modulus cannot be changed easily to alter the spring stiffness. However, as suggested for
variable spring stiffness in robotic actuators [11], the number of active coils could be changed. If

26
one of the spring supports is allowed to be threaded on the spring, the spring stiffness could be
changed by moving the support up and down the coils. A concept sketch can be seen in figure 4.

Figure 12. Sketch of adjustable straight coil concept with an upper support matching the coil pitch.

The upper spring support would be threaded onto the damper body to position it axially. It would
also have a helical support surface for the spring that matches the spring pitch which would allow
the spring to compress without bending. In such a way, the upper support could be positioned at
any part of the spring and thus inactivating the coils above the point at which the spring is
supported.

This was of achieving adjustable stiffness would be advantageous in terms of usability, cost,
manufacturing, and environmental robustness. The stiffness is adjusted through the position of the
upper support and a user does not need to keep track of any adjustment combinations. Adjustment
could also be made with the shock still attached to the bike (in most bike frames).

Material or manufacturing costs would not increase drastically compared to the coil shock
currently available. The same architecture and principal components would make up the shock and
new manufacturing techniques are not required. The spring itself is the same type as common
closed end springs. The open end could cause some issue in production when prestressing, but it
can also be manufactured as a closed end spring before having one end cut open.

The architecture of the shock also allows the damper to be separated from the spring and
completely sealed. Thus, the coil itself cannot cause any scratching or wear to functional sliding
surfaces and only one dynamic seal is open to the environment where contaminations can enter
the damper. A coil spring with relatively high pitch is also not very susceptible to trapping dirt and
abrasive particles that can cause premature spring failure due to wear. It is also easily protected
from corrosion by colour coating.

From eq. (7), the stiffness is inversely proportional to the number of active coils. Thus, doubling
the stiffness would require deactivation of half of the initial number of coils. With a straight coiled
spring, that would also mean halving the distance between the spring supports.

27
When the number of active coils is reduced, so is also the maximum stroke of the spring. An
adjustable spring must thus be designed to allow for the full shock stroke throughout its adjustment
range. The maximum spring travel is found by subtracting the bind height from the unloaded height
of the spring. The latter is found from the coil pitch, 𝛾, the number of active coils and the wire
diameter, 𝑑𝑡 :

𝐻0 = 𝛾𝑁𝑎 + 𝑑𝑡 + 𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 . (8)

The ends of the spring are inactive and have some stack height, 𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 , depending on the end
type. This does not change when the spring is compressed and does not affect the maximum stroke.
For the adjustable spring, one end would be open, and one end closed and ground to roughly 80%
of the wire diameter. At the open end, half a coil is inactive as it is the part supported. Thus, the
inactive height becomes:

𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 = 0,5𝛾 + 0,2𝑑𝑡 . (9)

At full travel, the spring is completely compressed, and the coils bind. The compressed height is
calculated using eq. (8), but with the wire diameter as the pitch:

𝐻𝑏 = 𝑑𝑡 (𝑁𝑎 + 1) + 𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 . (10)

Subtracting eq. (10) from (8) then gives the maximum stroke of the spring:

Δ𝐻max = 𝑁𝑎 (𝛾 − 𝑑𝑡 ). (11)

The maximum stroke is thus proportional to the number of active coils. In the stiffest and shortest
setting, the adjustable spring must have a stroke that is at least 1mm longer than the stroke of the
shock since the impact from coil bind would give an extremely harsh bottom-out. This would be
achieved by using a low total number of coils and a high pitch. However, a high pitch and low
number of active coils increases the stiffness as seen in eq. (7). To achieve the same stiffness with
a long stroke spring, the wire diameter must be reduced, or the mean diameter increased. This in
turns increases the material stress:

8𝑘Δ𝐻𝐷𝑚
𝜏= . (12)
𝜋𝑑𝑡3

A due to the curvature of the spring, stress concentration occurs in the wire [23] and a stress
concentration factor is therefore multiplied by the shear stress from eq. (12) to give a closer
estimate of the real maximum stress in the wire. This is referred to as the Wahl correction factor:
4𝐶−1 0,615
𝐾𝑤 = 4𝐶−4 + , (13)
𝐶

where 𝐶 is the spring indexed defined as the ratio of mean coil diameter to wire diameter:
𝐷𝑚
𝐶= . (14)
𝑑𝑡

The maximum stress in the wire is thus given by:

𝜏𝑚𝑎𝑥 = 𝐾𝑤 𝜏. (15)

28
In this way, designing an adjustable coil spring with a wide adjustment range becomes a non-linear
optimisation problem. The constraints are the stroke requirement and geometric constraints while
the objective function is to minimise the maximum stress in the spring.
A brute force search algorithm was used to find the best spring design for a given adjustment
range. This is outlined in the flowchart in figure 13.

Figure 13. Search algorithm for finding constrained optimum in adjustable coil spring design.

The stiffness and maximum shear stress were calculated for all combinations of design parameter
values within their specified ranges. If the stress was found to be lower than the previously
calculated minimum, the maximum travel of the spring at its stiffest setting was also calculated. If
this was greater than the shock stroke plus a margin of 1mm, the current parameter values were
saved as the latest optimum point and the calculated maximum stress was saved as a local

29
minimum point. If the travel of the spring did not exceed the minimum allowed value, the search
was continued, and the minimum point disregarded.

A longer spring generally allows for more travel and therefore a wider adjustment range.
Therefore, the free length of the spring was not selected as an optimisation variable. Instead, an
initial estimate of the maximum possible length was chosen based on the dimensions of the Intend
Hover air shock.

Figure 14. Estimated available space for spring and spring retainers.

Some stack heights would be required for the upper and lower spring retainers. A first estimate
was that this would be roughly 15mm and thus the free height of the spring was set as 180mm for
the concept design and optimisation.

Two steels were considered to give an estimate of the optimal spring dimension for the different
stiffness adjustment ranges under consideration. The first was music wire spring steel number
1.1200 with a shear modulus of 81,5GPa [24] and the second spring steel ASTM A227 Class I
spring steel with a shear modulus of 72GPa [25]. At this stage, the actual steel that would be used
for production was not known and thus the shear moduli were only used to estimate the required
dimensions and stress levels.

30
Table 4. Optimised spring parameters and estimated maximum stresses for two different spring steels.
Min Max Approx. Inner Wire Pitch Number Spring Max. Max.
Stiffness stiffness free diameter diameter [mm] of active travel at stress at stress at
[N/mm] [N/mm] height [mm] [mm] coils max. min. max.
[mm] stiffness stiffness stiffness
[mm] [MPa] [MPa]

ASTM A227 Class II: G=72000MPa, Rm≈ 1720MPa, Sy≈1030MPa


54,4 97,1 180 41,4 7,65 32,3 4,8 66,3 1230 2200
59,5 88,2 180 41,7 8,75 23,0 6,9 66,2 940 1390
64,6 83,3 180 41,6 9,3 20,2 7,9 66,5 870 1120
73,7 88,2 180 41,5 9,75 19,4 8,2 66,1 870 1040
59,3 74,3 180 41,9 9,25 19,2 8,3 66,3 810 1010
Music Wire EN 1.1200: G=81500MPa, Rm≈ 1380MPa, Sy≤920MPa
54,4 97,2 180 42 7,55 31,1 5 66,1 1270 2280
59,4 88,2 180 42 8,55 22,4 7,1 66,2 1000 1480
64,6 83,2 180 41,8 9,1 19,5 8,2 66,2 920 1180
73,6 88,1 180 42 9,55 19 8,4 66,3 930 1110
59,3 74,3 180 41,8 9 18,6 8,6 66,2 870 1090

The inner diameter of the coil was varied between 36mm and 42mm. A smaller diameter would
make it difficult to fit over a damper body and a larger diameter would likely cause issues with
frame compatibility. The number of active coils was varied between 4 and 10 while the wire
diameter was varied between 7mm-10mm.

Shear strength figures were taken from an online material database [25] for the A227 spring steel
and estimated from the ultimate tensile strength for the music wire steel. This estimate followed
rough guideline in machine design handbooks where it states that the elastic limit in torsion is 45-
60% of the ultimate tensile strength for music wire [23]. As these were very rough estimates for
allowable maximum shear stress, and the actual material for potential manufacturing was
unknown, the values were used only as a preliminary guideline to gauge whether the concept
would be feasible.

As seen in the table 4, it was found that a coil spring becomes severely over stressed if used to
give the ideal or very good adjustment range. Depending on material properties, the good and
acceptable ranges could well be achievable.

Coil spring with variable wire diameter


A development of the adjustable coil spring concept was to use a coil wire with varying cross
section diameter. As the stiffness is proportional to the wire diameter to the fourth power, a small
increase in wire diameter would cause a large increase in stiffness. If the open end of the coil has
a smaller wire thickness than the closed end, threading the upper support down the spring would
not only decrease the number of active coils, but also increase the effective wire diameter. The
stiffness increase would be non-linear with respect to the number of active coils, and the upper
support would have to be adjusted a shorter distance compared to on the straight coil spring for
the same stiffness range. This would allow the use of a shorter spring or potentially reduce the
maximum stress in the spring.

For parameter optimisation, the same constraints and search algorithm was used as for the straight
coil spring. However, the method for finding stiffness and stress had to be altered. The equations
used for calculations on the variable coil spring were combined from Shigley’s mechanical
handbook [23], and a modelling paper by Pop et Al. [26], where a mathematical model for stress
and deflection for a helical coil spring with variable diameter is proposed.

31
Assuming a coil spring is manufactured by winding the spring wire around a cylinder, the helical
spring will have a constant internal diameter. If the wire diameter is not constant however, the
coils do not stack directly on top of each other, and the minimum distance between each coil
depends on the wire diameter and the linear rate of change in diameter with respect to wire length.

Figure 15. Representation of coil spring with variable diameter at uncompressed state

In figure 15, an uncompressed spring with variable diameter is shown. In this state, the spring pitch
is the same between each coil and the wire diameter increases linearly from 𝑑𝑡0 to 𝑑𝑡,𝑒𝑛𝑑 :

𝑑𝑡,𝑒𝑛𝑑 −𝑑𝑡0 𝑑𝑟
𝑑𝑡,𝑛 = 𝑑𝑡0 + 𝑛 = 𝑑𝑡0 + 𝑑𝑁 𝑛, (16)
𝑁𝑎

where 𝑛 is the position of 0-𝑁𝑎 along the active spring helix. On the spring considered for this
concept, a closed and ground end would also be present at the end with the thickest wire diameter,
although this has been omitted for simplicity. The free height of the spring is calculated according
to eq. (8) but using the smallest and largest wire diameters:

𝐻0 = 𝛾𝑁𝑎 + 0,5(𝑑𝑡0 + 𝑑𝑡,𝑒𝑛𝑑 ) + 𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 . (17)

When the spring binds, the pitch is no longer the same for all coils as seen in figure 16.

Figure 16. Coil spring with variable diameter at bind height

32
If 𝑟𝑛 is the wire diameter at position 𝑛, then the minimum vertical distance between coil 𝑛 and coil
𝑛 + 1 is given by:

𝑑𝑟 2 𝑑𝑟 2
𝛾𝑚𝑖𝑛,𝑛 = √(2𝑟𝑛 + 𝑑𝑁) − (𝑑𝑁) . (18)

The bind height is then given by the sum of the minimum pitch between each coil:

𝐻𝑏𝑖𝑛𝑑 = ∑𝑁
𝑖 𝛾𝑚𝑖𝑛,𝑖 + 𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 (19)

This applies if 𝑁 is an integer, but if there is not a whole number of coils, then the approximate
bind height is:
𝑁
𝐻𝑏𝑖𝑛𝑑 ≈ ∑𝑖 𝑖𝑛𝑡 𝛾𝑚𝑖𝑛,𝑖 + 𝑁𝑑𝑒𝑐 𝛾𝑚𝑖𝑛,𝑒𝑛𝑑 + 𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 , (20)

where 𝑁𝑖𝑛𝑡 is the integer part and 𝑁𝑑𝑒𝑐 is the decimal part of 𝑁. The inactive stack height is
calculated according to eq. (9) with the largest wire thickness:

𝐻𝑖𝑛𝑎𝑐𝑡𝑖𝑣𝑒 = 0,5𝛾 + 0,2𝑑𝑡,𝑒𝑛𝑑 . (21)

Then the maximum spring stroke is given by the difference between the free height and the bind
height:

Δ𝐻𝑚𝑎𝑥 = 𝐻0 − 𝐻𝑏𝑖𝑛𝑑 . (22)

Since the wire diameter is not constant, eq. (7) cannot be used to calculate the spring stiffness for
a variable diameter helical spring. Instead, Castigliano’s theorem is utilised to find the spring force
over the spring stroke by taking the first partial derivative of the combined strain energy in the
spring [23] [26]:

𝐿 𝐷2 1
Φ = ∫0 𝑎 [ 4𝐽𝑚 + 𝐴] 𝑑𝑙, (23)

where 𝐽 is the wire cross section polar moment of inertia and 𝐴 is the area,

𝐺Δ𝐻
𝐹𝑠𝑝 = , (24)
Φ

𝐹𝑠𝑝,max
𝑘≈ . (25)
Δ𝐻𝑚𝑎𝑥

As the minimum pitch is not the same between each coil, the spring stiffness is not constant [26]
and eq. (25) gives only an estimate for the mean value. For the purpose of concept design, this was
assumed to be representative enough as long as the wire diameter change was not large.

For finding the length of the active wire, the top-down projection of the coil spring must first be
considered. This takes the form of an Archimedean spiral with the pitch:
𝑑𝑡,𝑒𝑛𝑑 −𝑑𝑡0
𝛼= , (26)
4𝑁𝑎 𝜋

and the start angle:


𝐷𝑚0
𝜃0 = , (27)
2𝛼

33
where 𝐷𝑚0 is the mean diameter at the spring end with the thinnest wire. For each coil, the change
in angle along the spiral is given by:

Δ𝜃 = 𝜃𝑛+1 − 𝜃𝑛 = 2𝜋Δ𝑁𝑛 . (28)

For each coil, Δ𝑁𝑛 =1, but for the last section, if the number of active coils is not an integer,
Δ𝑁𝑒𝑛𝑑 = 𝑁𝑑𝑒𝑐 . Then, the spiral length for each coil section is given by:
𝛼 2 2
𝐿𝑠𝑝𝑖𝑟𝑎𝑙,𝑛 = [𝜃𝑛+1 √𝜃𝑛+1 + 1 − 𝜃𝑛 √𝜃𝑛2 + 1 + ln (𝜃𝑛+1 + √𝜃𝑛+1 + 1) − ln(𝜃𝑛 + √𝜃𝑛2 + 1)] . (29)
2

The total active spring length is then:

𝐿𝑎 = ∑√𝐿2𝑠𝑝𝑖𝑟𝑎𝑙,𝑛 + 𝛾 2 . (30)
By using eq. (16)-(30) in the optimisation algorithm in figure 13 and adding the increase in wire
thickness as an optimisation parameter, estimates for maximum stress and optimal spring design
could then be made for variable thickness springs.

Table 5. Optimised spring parameters and estimated maximum stresses in a coil spring with variable wire
thickness for two different spring steels.
ASTM A227 Class II: G=72000MPa, Rm≈ 1720MPa, Sy≈1030MPa

Spring Max. Max.


Approx. Number travel at stress stress
Min Max free Inner Wire of max. at min. at max.
Stiffness stiffness height diameter diameter Pitch active stiffness stiffness stiffness
[N/mm] [N/mm] [mm] [mm] [mm] [mm/turn] coils [mm] [MPa] [MPa]
54 97,2 180 41,6 7,2-10,5 22,9 6,9 66,1 1420 1840
59,5 88,2 180 41,9 8,6-9,1 22,9 6,9 66,8 980 1390
64,6 83,4 180 40,2 9,1-9,25 19,9 8 66,1 890 1140
74,3 88,3 180 41,9 9,8-10 19,6 8,1 66,4 870 1020
59,6 74,4 180 41,4 9,2-9,25 19,6 8,2 67,1 820 1020
Music Wire EN 1.1200: G=81500MPa, Rm≈ 1380MPa, Sy≤920MPa
53,9 97,7 180 42 7,2-9,7 23,3 6,8 66,1 1430 1950
59,2 88,3 180 41,7 8,5-8,55 22,7 7 66,1 1000 1490
64,4 83,3 180 41,7 9,1-9,15 19,7 8,1 66,2 910 1180
73,1 88,3 180 41,1 9,4-9,45 19,2 8,3 67,1 940 1140
59,4 74,5 180 41,9 9-9,1 18,8 8,5 66,1 870 1080

Torsion Spring
A thin-walled metal tube acts as simple torsion springs when twisted. The torque, 𝑇, required to
cause an angular deflection, Ω, of an isotropic thin-walled tube or rod is given by:

𝐺𝐽Ω
𝑇(Ω) = , (31)
𝐿

where 𝐺 is the shear modulus of the material, 𝐽 is the second polar moment of area of the cross
section and 𝐿 is the length of the tube or rod [27]. Thus, the torsional stiffness is inversely
proportional to the active length of this type of torsion spring.

The torsion spring concept was based on a thin-walled circular tube where the active length could
be adjusted by having several clamping positions at one end. A high pitch lead- or ball screw

34
would then convert an axial force to a twisting moment. This would allow the torsion bar to be
used as a linear spring for compatibility with current bike frames.

Figure 17. Sketch of torsion spring concept implemented in an MTB shock.

The main advantages of this concept were ease of adjustability and fine tuning of the spring
stiffness. In order the change the stiffness, the user would simply change the clamping position at
one end of the spring. This would also allow for stepless adjustment making it possible to tune the
spring exactly to each rider.

Several issues arose early in the development of this concept when adapting it to current bike
frames and the shock size standard considered in this project. The most critical issue identified
was that the mounting points on current bike frames are not designed to take up a twisting load
which would be necessary with this type of spring. Therefore, a torsion spring could potentially
cause failure of some bike frames.

Another issue would be the contact between the thread and the nut parts of the spring. If designed
with a sliding contact, friction would be difficult to predict and the damping more difficult to tune.
If designed with a rolling contact, like a ball screw, then the spring would require a multitude of
moving parts which adds complexity and make service difficult. In addition, the use of a thread on
the torsion bar would make the system difficult to seal from contaminations.

Although allowing for simple stiffness adjustment, this concept was abandoned due to poor
compatibility with current bike frames.

Coil/air hybrid spring


One concept idea was to improve the sensitivity of a traditional air spring by combining it with a
coil spring in series. This would allow the stiction free coil spring to compress before the break
away force of the air spring is overcome. This type of spring can be found in suspension forks
from Extreme Racing Shox and DT Swiss [28] [29].

The air spring would consist of two chambers - like most current air sprung shocks. An external
valve allows for pressure adjustment in the positive chamber and an equalisation port would
pressurise a negative chamber. The pressure in the negative chamber would counteract the preload
from the positive chamber so that seal friction would be the sole cause for the required break-away
force.

The main advantage of this concept is increased sensitivity compared to common air springs and
increased ramp up compared to common coil springs. As the serial coil spring binds, the air spring

35
with its inherently progressive spring curve takes over. Thus, a lower initial spring rate could be
used without risking harsh bottom-outs.

A severe disadvantage is the space requirement. The spring curve of an air spring with dual
chambers features three sections that are more or less distinct depending on the chambers’ volume
ratio. Assuming zero seal friction, the stiffness in the initial part is relatively high but is drastically
reduced for the middle portion of the spring travel. Towards the end of the stroke, the spring
becomes increasingly stiffer depending on the compression ratio.

Assuming a closed system, following the ideal gas law, the pressures in each air chamber are given
by:

𝑝0 𝑉0 = 𝑝1 𝑉1, (32)

where 𝑉0 is the initial volume of the air chamber and 𝑝0 is the initial pressure. At rest, the forces
on either side of the air piston are balanced and no net force is exerted on the spring shaft. When
the spring is compressed, the volume in the positive chamber decreases while the pressure
increases, and vice versa in the negative chamber. Thus, a net force in the opposite direction of
compression is generated and increases over the stroke of the shock. Depending on the relative
sizes of the positive and negative chambers, the rates of change in piston force with regards to
spring position will be different. Therefore, the shape of the initial parts of the spring curve will
depend heavily on the relative sizes of the two air chambers. This is displayed in figure 18.

Figure 18. Idealised spring force curves for dual chamber air springs with the equal pressures and
different chamber size ratios.
When the negative chamber is small compared to the positive, the spring is significantly stiffer at
the initial portion of travel than the middle portion. This is seen in the blue spring curve in figure
18. When the chambers have the same initial volume (purple curve), the spring stiffness increases
gradually over the entire stroke. A sharp increase in stiffness towards the end stroke is displayed
by all chamber ratios and this is dependent on final compression ratio of the positive chamber.

36
A falling spring rate, as generated by a low negative volume ratio, is undesirable as it creates a
harsh ride feel at the initial part of the shock travel. It also causes the shock to sit too deep in its
travel as the middle portion of the spring stiffness is relatively low. Therefore, it is desirable to
maximise the volume of the negative chamber. From figure 18 it is seen that the volume of the
negative chamber should be somewhere in the range of 30-100% of the positive chamber to give
a supportive spring curve.

Excessive progression becomes an issue if the Compression Ratio (CR) of an air spring is too
great. If the spring becomes too stiff at the end of its travel, the bike rider experiences a very harsh
end stroke and may struggle to maintain control of the bike after hard impacts. Some progression
is desirable though. This allows the rider to make use of a larger portion of the spring travel while
reducing the risk to blow through the entire travel and experience harsh bottom outs.

Figure 19. Dual chamber air spring curves for a 40% chamber volume ratio, same pressure, and
different compression ratios.
For a stroke of 65mm, an air spring with a maximum CR of 3 would require roughly 87mm of
positive air chamber length. With a 40% volume ratio between the chambers, the total length of
the air spring would have to be about 122mm. Adding an approximate spring piston thickness of
roughly 5mm leaves about 70mm for a mechanical spring.

The concept idea was to keep the architecture of a traditional coil spring damper and replace the
upper spring retainer with an air sleeve where an air piston would be allowed to move. This piston
would then act as the upper spring support and a shorter coil spring with two closed and ground
ends would connect the piston to the damper shaft. A sketch can be seen in figure 20.

37
Figure 20. Concept sketch of air/coil hybrid spring

The first identified issue with this design, as can be seen in figure 20, is that there is no easy way
to include a negative air chamber. If this was to be fitted, two additional dynamic seals would be
required, and the negative chamber volume would become very small. Also, with the estimated
length requirements for a good air spring curve, the length of the air sleeve would likely prevent
the shock from using full travel. Therefore, the concept was adapted to using a single air chamber
activated directly by the shorter coil spring.

This would give the spring two or three distinct sections in the spring curve depending on the
stiffness of the coil spring and the initial pressure in the air spring. A shoulder would prevent the
air piston from coming out the bottom of the air sleeve. Thus, when pumped up, the air spring
would be preloaded. This, together with the air pressure would allow the hybrid spring to be
adjusted and the coil spring would have a set stiffness. When the shock goes through its travel,
only the coil spring would compress until the spring force equals the preload force on the air piston.
When this threshold force is reached, the air spring is activated and both springs compress.
Depending on the air pressure and the coil spring design, the coil spring may bind at some point
throughout the shock stroke and the air spring would allow for full shock travel alone.

38
Figure 21. Spring curves for different configurations of the air/coil hybrid spring concept.

Some initial spring curves were generated for this type of hybrid spring to have an idea of the
spring characteristics. This is shown in figure 21 where the combined forces from combinations
of 2 different coils spring stiffnesses and initial air pressures were plotted over the total shock
stroke. The maximum stroke of the coil spring was set to 25% of the total shock stroke. A major
drawback was made evident here as the spring curves are very digressive. Thus, such a spring
would give a harsh initial stroke and ride very deep in its travel which is very undesirable. Also, if
the air spring is preloaded too heavily, the coil spring can bind before the air spring initiates travel
which would cause a shock load on the system and make the system extremely harsh.

Another drawback would be the risk of contamination of the sliding surfaces for the air piston and
risk of abrasion from the coil spring. Such a design would require wipers and it is possible to have
sliding surfaces exposed to the environment without extreme wear. An example can be found in
suspension forks for mountain bikes. However, such exposure would require shorter service
intervals which is not desirable.

A wide range of adjustability could thus potentially be achieved by an air/coil hybrid spring, but
the adaptation to MTB would yield a poor spring curve and drawbacks in robustness to
environmental exposure.

39
3.4 Concept evaluation
Qualitative evaluation
During the concept generation stage, a qualitative assessment of each concept regarding suitability
for mountain bikes was performed before advancing to more detailed analysis. Since there were
no detailed concept designs or sketches at this stage, estimates of concept feasibility were made
by answering questions in the following categories:

• Frame compatibility – Could the spring be adapted for shocks fitting current MTB frames?
• Environmental robustness – Could the spring perform in all types of riding conditions
without extreme wear or damage?
• Usability – Would a rider be able to easily adjust spring stiffness?
• Manufacturability – Could the spring components be easily manufactured?

Adjustment range was not included in this stage of evaluation since such detailed analysis had not
yet been completed. Each concept had been developed based on the theoretical adjustability and
were therefore assumed to achieve some level of adjustability. Performing such analysis was
however deemed futile for a concept that could not be adapted for application in MTB shocks.
Therefore, if it was the answer was a strict no in either of the assessment categories, a concept
would be discontinued before the detailed analysis stage. The result from this qualitative
evaluation is presented in table 6.
Table 6. Results from qualitative concept evaluation of the four initial adjustable spring concepts.
Straight coil Coil spring Air/Coil hybrid Torsion spring
spring with variable spring
wire thickness
Yes: Can likely Yes: Can likely Yes: Can likely be No: Would
be fitted to a be fitted to a fitted to a shock generate a torsion
shock like current shock like current similar to current load on shock
Frame coil springs coil springs coil spring shocks mounts which
Compatibility they are not
designed for.

Yes: Requires Yes: Requires Maybe: Requires Maybe: Could


one exposed one exposed at least 3 exposed potentially be
dynamic seal. dynamic seal. dynamic seals and designed with one
One exposed One exposed two exposed exposed dynamic
Environmental surface that could surface that could surfaces that could seal and sliding
robustness be damaged by be damaged by be damaged by surface. Difficult
abrasive abrasive abrasive particles. to seal thread
particles. particles.

Yes: Would Yes: Would Yes: Would require Yes: Would


require adjusting require adjusting adjustment of air require
Usability upper spring upper spring pressure through repositioning of
retainer. retainer. accessible valve. clamp.

Yes: Could be Maybe: Would Yes: Would require Yes: Would be


made from require tapered common extruded made from
common spring spring wire and aluminium and extruded
wire like common manufactured like manufactured by aluminium and
Manufacturability springs. common springs. turning or milling. could be
manufactured by
turning or milling.

40
Out of the four concepts, only the torsion spring concept did not pass the qualitative assessment
due to incompatibility with current mountain bike frames. That is not to say there is no frame that
could potentially withstand torsion loads from the shock, but since no frames are specifically
designed for such loads, it is not safe to design a shock that generates such loads. The only concept
that clearly passed the assessment was the straight coil spring. A coil spring with variable wire
diameter may cause problems in production and an air/coil hybrid spring may be less resistant to
wear in harsh riding conditions. Neither of these drawbacks were cause for rejection however and
they thus continued to detailed analysis.

Quantitative Evaluation
For the final choice of concept, desirability scoring was used. This is a quantitative evaluation
method that allows different metrics, even qualitative, to be compared in a clear way with
quantitative scores. Based on targets and limits, each concept receives a normalised score between
zero and one in each evaluation category. The closer to one, the closer to the target a concept is in
a category.

Finding the overall best scoring concept is to combine the scores from each category and while
doing so, the importance of each category can be included. In this way, the best concept can clearly
be established by either comparing the category scores or the overall best score. Also, if a concept
is not within the limits in a category, it is likely not a viable alternative and the overall score
becomes zero.

In each category, a target is defined along with one or two limits. If the direction of improvement
is lower-is-better, then an upper limit is set. If the direction is higher-is-better, then a minimum
limit is set and for a target-is-best case, both upper and lower limits are set. Then the category
score is calculated according to the following relations [30]:

Highest-is-best:
0, 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑙𝑜𝑤𝑒𝑟
𝑎𝑐𝑡𝑢𝑎𝑙−𝑙𝑜𝑤𝑒𝑟 𝑤𝑒𝑖𝑔ℎ𝑡
𝑑𝑠𝑐𝑜𝑟𝑒 = {( ) , 𝑙𝑜𝑤𝑒𝑟 ≤ 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑡𝑎𝑟𝑔𝑒𝑡 } (33)
𝑡𝑎𝑟𝑔𝑒𝑡−𝑙𝑜𝑤𝑒𝑟
1, 𝑎𝑐𝑡𝑢𝑎𝑙 ≥ 𝑡𝑎𝑟𝑔𝑒𝑡
Lowest-is-best:
0, 𝑎𝑐𝑡𝑢𝑎𝑙 ≥ 𝑢𝑝𝑝𝑒𝑟
𝑢𝑝𝑝𝑒𝑟−𝑎𝑐𝑡𝑢𝑎𝑙 𝑤𝑒𝑖𝑔ℎ𝑡
𝑑𝑠𝑐𝑜𝑟𝑒 = {( ) , 𝑡𝑎𝑟𝑔𝑒𝑡 ≤ 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑢𝑝𝑝𝑒𝑟 } (34)
𝑢𝑝𝑝𝑒𝑟−𝑡𝑎𝑟𝑔𝑒𝑡
1, 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑡𝑎𝑟𝑔𝑒𝑡
Target-is-best:
0, 𝑎𝑐𝑡𝑢𝑎𝑙 ≥ 𝑢𝑝𝑝𝑒𝑟 𝑜𝑟 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑙𝑜𝑤𝑒𝑟
𝑢𝑝𝑝𝑒𝑟−𝑎𝑐𝑡𝑢𝑎𝑙 𝑤𝑒𝑖𝑔ℎ𝑡
𝑑𝑠𝑐𝑜𝑟𝑒 = (𝑢𝑝𝑝𝑒𝑟−𝑡𝑎𝑟𝑔𝑒𝑡) , 𝑡𝑎𝑟𝑔𝑒𝑡 ≤ 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑢𝑝𝑝𝑒𝑟 . (35)
𝑎𝑐𝑡𝑢𝑎𝑙−𝑙𝑜𝑤𝑒𝑟 𝑤𝑒𝑖𝑔ℎ𝑡
{(𝑡𝑎𝑟𝑔𝑒𝑡−𝑙𝑜𝑤𝑒𝑟) , 𝑙𝑜𝑤𝑒𝑟 ≤ 𝑎𝑐𝑡𝑢𝑎𝑙 ≤ 𝑡𝑎𝑟𝑔𝑒𝑡 }

By assigning a weight factor to the category score, the scoring function becomes non-linear. With
a weight lower than one, a deviation from the target does not generate as large a score reduction
while the opposite is true for a weight greater than one.

41
When combined to an overall score, the following formula is used [30]:
1
𝑖 𝑞
[∏𝑞𝑗=1 (𝑑𝑗 𝑗 ) ] 𝑗=1 𝑗
∑ 𝑖
𝐷𝑠𝑦𝑠𝑡𝑒𝑚 = . (36)

Qualitative categories were given a numerical scoring system based on order of quality and a target
value equal to the number corresponding to the highest quality. Quantitative categories were
scored directly on concept numerical values. Table 7 lists the categories used and outlines of the
scoring system in each category while table 8 lists the targets, limits, directions of improvement,
weights and importance values used for concept scoring
Table 7. Explanation of evaluation categories and outline of scoring systems
Evaluation Category Description and scores
Usability - Number of dials The number of dials or features that must be controlled to set
the stiffness

Usability - Accessibility of Additional operations to adjust stiffness: 0=none, 1=loosening


adjustment one or two screws, 2=dismount one end of the shock,
3=dismount entire shock, 4=disassemble shock/spring
Manufacturability Manufacturing effort: 0=only standard parts, 1=custom parts with
standard base materials and common manufacturing methods,
2=custom parts with custom base materials and common
manufacturing techniques, 3=custom parts with custom base
materials and specialised manufacturing techniques

Spring Characteristics Shape of spring curve: 4=slightly progressive, 3=linear,


2=smooth digressive, 1=non-linear with sharp transitions
between sections
Adjustment range Based on pre-study: 5=Ideal range, 4=Very good range,
3=Good range, 2=Either of the acceptable ranges, 1=less than
either acceptable range

Material Stress Fraction of yield stress eg. 0-1


Environmental robustness Number of dynamic seals open to environment
- Exposed dynamic seals

Frame compatibility Compatibility with current bike frame shock mounts. 1=yes,
0=no

42
Table 8. Targets, limits, directions of improvement, weight and importance values for each category used
in the concept desirability scoring
Direction of
Target Limit improvement Weight Importance

Usability - Number of dials 1 2 Lowest-is-best 1 1

Usability - Accessibility of
adjustment 0 4 Lowest-is-best 1 1
Manufacturability 0 3 Lowest-is-best 2 2
Spring Characteristics 4 1 Highest-is-best 1 1
Adjustment range 5 1 Highest-is-best 1 2

Material Stress 0,5 1 Lowest-is-best 2 1


Environmental robustness
- Exposed dynamic seals 0 4 Lowest-is-best 1 1

Frame compatibility 1 0 Highest-is-best 1 2

Based on the scoring template in table 7, each concept was given numerical scores in each category
and the desirability scores were calculated using equation (33)-(36). After it was established that
the air/coil hybrid concept exhibited an undesirable spring curve, detailed analysis was stopped,
and stress or adjustment range were not established. Still, all concepts were given full desirability
scores to give an overall comparison, but in some categories, the torsion spring and air/coil hybrid
concepts were given very roughly estimated scores. This allowed easier comparison in each
category while still allowing the underperforming concepts to be discarded in the final choice as
their overall score would be zero. The numerical scores and desirability scores are presented in
table 9 and 10.

43
Table 9. Numerical category scores for each concept

Manufacturability
Number of dials

Accessibility of

Characteristics

Material Stress

Environmental

dynamic seals

compatibility
robustness -
Adjustment
adjustment
Usability -

Usability -

Exposed
Spring

Frame
range
1: 1: Spring 1: Custom 2: Concept 0,98: 1: Can
dimensions, design Max stress 1: One
Straight Position fastening easily be
standard wire, 3: Linear shows reaches dynamic seal
of upper screw must designed for
coil spring spring be
manufactured acceptable 98% of open to
current
like common range can yield environment
retainer untightened frames
springs be achieved strength

2: Custom 2: Concept 0,99:


Increasing 1: 1: Spring dimensions,
1: One 1: Can
design Max stress
wire Position fastening custom tapered 4: Slightly easily be
shows reaches dynamic seal
of upper screw must wire, designed for
thickness spring be manufactured
progressive acceptable 99% of open to
current
coil spring range can yield environment
retainer untightened as common frames
be achieved strength
springs

1: Likely to be
1: Air manufactured 1: Digressive 3*: 0,5*: 3: Minimum of 1: Can
Air/coil pressure 0: No from common with sharp easily be
hybrid Abandoned Abandone 3 dynamic
adjustme additional materials using transitions designed for
before d before seals open to
spring nt through operation common between current
analysis analysis environment
1 valve milling/lathe sections frames
technique

1: 1: 1*: 3*: 0,5*: 1*: 1: Not


Torsion Loosening Abandoned Abandone Abandoned compatible
Position Abandoned 3: Linear
Spring of clamp
of clamp
before analysis
before d before before due to
mechanism analysis analysis analysis torsion load

Table 10. Weighted desirability scores and overall system scores for each concept.

Frame compatibility
Usability - Number

Overall Combined
Adjustment range

Exposed dynamic
Manufacturability
Accessibility of

Characteristics

Material Stress

Environmental
robustness -
adjustment
Usability -
of dials

Spring

Score
seals

Straight coil
1 0,75 0,44 0,67 0,25 0,0016 0,75 1 0,342
spring

Increasing
wire
1 0,75 0,11 1 0,25 0,0004 0,75 1 0,243
thickness
coil spring

Air/coil
hybrid 1 1 0,44 0 0,5* 1* 0,25 1 0
spring

Torsion
1 0,75 0,44* 0,67 0,5* 1* 0,75* 0 0
Spring

44
Choice of Concept
The final choice of concept was based on the overall system desirability score. Although a concept
would be superior in one category, if not reaching the minimum requirement in any other category
it cannot be a good end-product. Therefore, the overall score had to be the final decision base.

Seeing that the straight coil spring concept scored the highest overall, this was chosen as the final
concept. From the category scores however, it could be seen that it was not the clear winner across
the field. The variable wire thickness coil spring exhibited better spring characteristics with a
slightly progressive spring curve while the air/coil hybrid spring would require less operations to
adjust. Also, the stress level in the straight coil spring was very close to the material yield strength
based on the concept analysis. The same was true for the variable wire thickness coil spring. The
low score in this category was the cause for the relatively low overall score. However, the material
data used was just used as estimates and the analysis showed that it would likely be possible to
achieve close to an acceptable adjustment range without over stressing the spring. After contacting
spring manufacturers and obtaining real material data, the design would likely change, and the
final stress levels could be managed. Therefore, it was decided to continue to the detailed spring
design and shock design stage with the straight coil spring concept.

3.5 Detailed design


Shock and damper construction
Based on existing parts at Intend BC, it was decided that the first prototype shock would follow a
recirculating damper architecture. This would allow the use of already manufactured damper tubes
with hard anodised and polished inner surfaces which would reduce parts manufacturing cost and
time. The main damper shaft and lower eyelet from the Intend Hover shock would also be used
along with parts for an IFP assembly. To be designed and manufactured thus remained the upper
shock body, valve piston assembly, damper piston, outer shock tube, upper- and lower spring
retainer, and seal head. In figure 22, the locations of the parts are shown on an early version of the
shock design.

Figure 22. Section views of an early version of the recirculating damper design with components marked
out.

45
The first step in the shock design was to establish the available space for the coil spring before the
detailed spring design could be completed. This was done by making a first rough CAD model of
the shock with necessary seal glands and threads and finding the maximum distance between the
lower spring retainer and the shock body.

The outer tube would be threaded on the outside with an M33x1 thread for interfacing with the
shock body and the upper spring retainer. This is the same thread used on RockShox coil shocks
which would allow interchangeability of springs. With a thread engagement of 6 threads in the
shock body, that left around 179mm where the spring and upper spring retainer should fit.

Figure 23. Available space for adjustable coil spring assembly.

A solid 22mm damper piston was chosen based on the available anodised tubes. Smaller tubes
with an inner diameter of 16mm were also available, but a larger piston was deemed more suitable
for generating enough damper force.

The damper piston would displace oil in the inner tube, through the shock body and then through
the compression valve. An oil reservoir pressurised by an IFP would allow oil to flow through a
check valve in the in the rebound valve assembly, through the space between the inner and out
tubes and then back into the inner tube behind the damper piston. In the rebound stroke, the flow
would the reverse. The IFP would allow for oil volume compensation due to oil displacement by
the damper shaft and would be pressurised through a Shrader valve by a common shock pump.

Linear guiding of the damper piston and damper shaft would be handled by a bushing pressed into
the seal head and a PTFE glide ring on the damper piston.

An o-ring would seal the shaft and a scraper would prevent contaminants from entering the
damper. Dynamic o-ring seals were designed with 10-15% compression for minimal friction while
static seals were designed with 20-25% compression for maximum sealing.

Damper modelling
Two types of valve pistons were available: one with an outside diameter of 16mm and one with
23mm. Both featured three large ports for main flow and three smaller ports for check flow. The
pistons are shown in figure 24

46
Figure 24. Valve piston alternatives and their port sizes.
Rough modelling of damper forces was made to determine which piston would be chosen for the
shock. A small piston would allow a sleeker shock with smaller overall dimensions. However, if
the ports are too small, there is risk for cavitation and over-dampening at high shaft velocities. A
larger piston on the other hand would allow a wider range for tuning with shims since the ports do
not cause excessive flow restriction at normal shaft speeds.

The model used was a modification of a numerical damper model for monotube shocks presented
by Rhoades [9]. His model is non-dynamic and relates the pressure difference over each valve and
port in the damper to the flow rates due to different damper piston velocities. This gives a system
of linked non-linear equations that can be solved numerically to give the total damper force and
chamber pressures over a range of shaft speeds.

In the model, the valve shims are modelled as linear springs so that the valve opening is
proportional to the pressure difference across it. The spring stiffness is found by FEM modelling
of the shim or shim stack. A pressure is applied to one face of the shim over an area the same shape
as the piston ports while a central portion of the shim, the size of a clamp shim, is fixed. The
resulting deflection of the shim edge is then assumed to be the valve opening and the stiffness is
estimated as the pressure divided by the deflection.

Since the model proposed by Rhoades is given for a monotube damper, additional flow equations
had to be added to adapt it for the recirculating damper designed in this project. In the recirculating
damper, additional ports exist between the compression chamber and the compression valve, the
compression valve and the oil reservoir, and the rebound valve and the rebound chamber. These
were all modelled as circular ports since they would be drilled in the shock body. In addition, the
solid piston in the recirculating damper would prevent flow across the damper piston in contrast
to the monotube design. Figure 25 Shows the flow path in the damper.

47
Figure 25. Flow through additional channels and valves in the recirculating damper.

The channel linking the compression chamber to the compression valve had a diameter of 7mm
while the rebound channel was 5mm. Larger dimensions would have been desirable but was not
possible if a safe wall thickness in the shock body was to be obtained. For the modelling, each
valve was given a 0.1𝑚𝑚 thick shim without preload and the bleed port was fully open with a total
flow area of 6,8𝑚𝑚2. The IFP pressure was set to 100psi or roughly 690𝑘𝑃𝑎. In reality, the IFP
pressure increases throughout the stroke as the air chamber volume is reduced. However, to have
conservative estimates, the pressure was assumed to remain constant. The damper oil viscosity
was set as 881𝑘𝑔/𝑚3 with a kinematic viscosity of 20𝑐𝑆𝑡. These were values for Motorex Racing
Fork Fluid 2.5wt [31] that would be used. Damper pressures and forces estimated by the model
are presented in figure 26 and 27.

Figure 26. a) Chamber pressures in the damper during compression and rebound with small pistons and
100psi IFP pressure. b) Damping force over shaft speeds with small piston and 100psi IFP pressure.

48
Figure 27. a) Chamber pressures in the damper during compression and rebound with small pistons and
100psi IFP pressure. b) Damping force over shaft speeds with small piston and 100psi IFP pressure.

The smaller piston generates roughly twice the damping force of the large piston, and the chamber
pressures reach negative values. Negative pressures cannot exist, so the plotted values are not
accurate beyond the zero crossings. However, the negative pressures show that cavitation occurs.
This happens only with the small piston and at relatively high, but not extreme shaft speeds.
Cavitation can be avoided by increasing the IFP pressure, but this also reduces the sensitivity of
the shock since it adds a preloading force on the damper shaft.

The increased damper forces and cavitation from the small piston showed that its ports generated
excessive flow restriction. As damper force can always be increased with stiffer shim valving, the
large piston would therefore be the best choice by allowing a wider damper tuning range. It would
also allow lower IFP pressure which would increase the overall sensitivity. Therefore, based on
the damper model, the large piston was chosen for the damper valves.

Spring adjustment feature


Spring stiffness adjustment would be achieved by allowing the upper spring retainer to thread up
and down the spring wire. By doing so, the section of the spring above the retainer would be
deactivated and the stiffness would increase. As an increased free spring height would increase the
maximum travel of the spring and reduce the maximum stress, the spring retainer would preferably
have as low stack height as possible.

The minimum stack height was established by FEM analysis in Solid Edge. From the concept
design stage, it was established that the maximum achievable stiffness would be just under
90𝑁/𝑚𝑚, a maximum load of 5,9𝑘𝑁 was used as this would be just over the maximum expected
spring force at 65mm travel. The upper retainer was fixed on its inner cylindrical face to represent
the threaded connection to the outer damper tube. The load was then applied on the spring
interfaces as an evenly distributed force in the upwards direction. Dimensions were iteratively
reduced until a safety factor of 2 had been reached. For the simulation, the built-in material library
was used to define the material as 7075-T6 aluminium with a Young’s modulus of 71,7𝑀𝑃𝑎 and
a Poisson’s ratio of 0,33. The resulting von Mises stresses in the final model are shown in figure
28 and 29.

49
Figure 28. von Mises stresses in upper spring retainer from finite element simulation on final version.

Figure 29. Back view of simulation results with maximum stress area visible
One critical area was identified where the stress reached roughly 204𝑀𝑃𝑎. As it would be
manufactured from 7075-T6 aluminium with a yield stress of 470𝑀𝑃𝑎 [27], this would give an
overall minimum safety factor of 2,3 for yield.

Tetrahedral elements were used for the simulation and less than 0,5% difference in maximum
stress between 0,77mm elements and 0,45mm element showed that the results were mesh
independent.

The general dimensions derived by this dimensioning strategy are shown in figure 30.

50
Figure 30. Minimum general dimension for upper spring retainer to give enough safety for yield
Based on these dimensions, the maximum free height of the spring could be established, and the
spring parameters optimised. The pitch and pitch diameter of the spring interface could then be
altered somewhat to accommodate the optimised spring design without risk of material yield at
maximum travel if the established dimensions were not reduced.

As the high coil pitch would cause the thread-like interface between the spring retainer and the
spring to not be self-locking, a securing screw was added to the retainer. This would consist of a
standard countersunk head screw that presses the spring wire against the retainer. The angled head
of the screw would contact the spring, creating a force in both the axial and radial direction of the
damper tube. Thus, an increased frictional force between the retainer and spring would prevent
relative movement between the components which would prevent them from losing contact during
operation.

Optimisation of final spring design


Before spring optimisation, spring manufacturers were contacted by the industrial supervisor. A
positive response was received from Force Technology. This is a UK based spring manufacturer,
and their sales representatives and engineers were willing to manufacture a small batch of
prototype coil springs and assist with spring design for the project. Due to the Covid-19 pandemic,
base material was scarce though. Only 8,75𝑚𝑚 wire from OTEVA 70 SC steel [32] was readily
available in stock. The delivery times for other materials was highly uncertain and were likely in
the order of 4-8 months. Optimisation of the spring was therefore made with the available wire
diameter and the material data for OTEVA 70 SC spring steel.

The steel type is produced by Suzuki-Garphyttan and the nearest equivalent steel grades are EN
VDSiCr or SIS 142090-05 [32]. It is a steel mainly used for valve- clutch- and transmission springs
and offer extremely high fatigue properties. The specified Young’s modulus and shear modulus
are 206𝑀𝑃𝑎 and 79,5𝑀𝑃𝑎 respectively [32].

A maximum free spring height of 173𝑚𝑚 was used for the optimisation to leave some margin for
removal of the lower spring retainer and spring adjustment. The spring design was then optimised

51
according to the method used in the concept generation. In table 11, the final optimised spring
parameters are presented.

Table 11. Final optimised spring parameters for different stiffness ranges using OTEVA 70 SC spring
steel.
OTEVA 70 SC: G=79500MPa, Rm=1660-1760MPa
𝚫𝑯𝒎𝒂𝒙 𝝉𝒎𝒂𝒙 𝝉𝒎𝒂𝒙 𝝉𝒎𝒂𝒙 𝝉𝒎𝒂𝒙
𝒌𝒎𝒊𝒏 𝒌𝒎𝒂𝒙 𝑯𝟎 𝑫𝒊 𝒅𝒕 𝜸 @ @ @
𝑵𝒂 𝑹𝒎,𝒎𝒊𝒏 𝑹𝒎,𝒎𝒂𝒙
[mm] [N/mm] [mm] [mm] [mm] [m] 𝒌𝒎𝒂𝒙 𝒌𝒎𝒊𝒏 𝒌𝒎𝒂𝒙 [%] [%]
[mm] [MPa] [MPa]

54 98 173 Insufficient spring travel at max stiffness for any configuration to reach
59 89 173 adjustment range
64,6 83,2 173 40,2 8,75 19,9 7,7 66,5 990 1280 77,1 56,3
73,1 88,1 173 38 8,75 19,6 7,8 70,4 1090 1310 78,9 61,9
59,5 74,3 173 40,7 8,75 18,9 8,1 66,2 920 1150 69,3 52,3
70,4 87,2 173 38 8,75 19 8,1 66,7 1050 1300 78,3 59,7

The supporting engineer from the spring manufacturer stated that the common design rule for
springs was that the corrected maximum shear stress should not exceed 70% of the material’s
tensile strength. For the specified material, the tensile strength was 1660-1760 MPa [32]. Based
on this data, a spring allowing for a range of roughly 59-74 N/mm would be withing the 70%-limit
within the entire range in variability in material tensile strength. This is the configuration
corresponding to the green row in table 11. If only considering the upper limit of material strength
however, all configurations would be within the 70%-limit. For the stiffness ranges 54-98 N/mm
and 59-89 N/mm, the dimensional constraints did not allow sufficient spring travel over the entire
adjustment ranges.

Based purely on the stress data from the optimisation, a 59-74N/mm should have been selected for
the first set of prototypes. However, since the spring would have to be tested, it was necessary to
consider weights of the potential test riders. These were mostly men weighing between 70-95kg
and therefore the lighter spring would not be suitable. The second-best option in terms of stress, a
65-83N/mm spring, would also not cover the heavier test riders. Due to this, a spring for the range
of 70-87N/mm was selected. This is not strictly within one of the target ranges established in the
pre-study, but it is close to the upper acceptable range of 74-89N/mm and it offers a wider span.
Also, the maximum stress levels estimated occur at maximum shock travel which should not occur
frequently in real use. Therefore, it would also be desirable to field test a potentially slightly over
stressed spring to establish whether the expected low number of maximum stress cycles allows an
over stressed design.

Thus, the manufacturing specifications for the spring were the following:
• Wire diameter: 8.75mm
• Inner diameter: 38mm +0.4/-0.0mm
• Number of active coils: 8.6 ±0.1mm
• Number of coils at closed and ground end: 1
• Total number of coils: 9.6 ±𝟎. 𝟏
• Ends: 1 closed and ground, 1 open
• Pitch: 19mm ±0.1mm

The weight of this spring was estimated to 672 g by the manufacturer.

52
Final CAD assembly
The final spring and shock designs are presented as CAD assemblies and an assembly drawing
with general dimensions in figure 31-34.

Figure 31. Final shock and spring assembly with spring adjusted to a) minimum stiffness=70N/mm and b)
maximum stiffness=87N/mm.

Figure 32. Section view of final design.

53
Figure 33. Section view of final damper valving and flow channels.

Figure 34. Assembly drawing of final shock and spring assembly with general dimensions and shock
stroke.

54
3.6 Test strategy
Testing of the spring would serve two main purposes. The first would be to physically verify the
models used for design. The second would be to establish the actual limit for allowable theoretical
maximum material stress in the spring.

Model verification
Model verification would be done by quasi static testing of the shock. By measuring the force over
the shock stroke for multiple spring stiffness settings, the actual achievable spring stiffness range
can be established. The following test method is proposed:

• Mount shock with spring and open IFP valve in an axial tensile test machine.
• Set the spring stiffness to the minimum setting.
• Compress the shock slowly (<2mm/s) for at least 60mm and measure the spring force
throughout the stroke.
• Repeat with the spring setting at 25%, 50%, 75% and 100% of maximum stiffness.

For an estimate in the variability, this test should then be performed with at least 3 prototype
springs.

Establishing effects of spring stress in MTB application


When a spring is over stressed, it collapses, and the stiffness is reduced to zero beyond the critical
stroke length. Plastic deformation occurs and the spring can set causing reduction in free height
and pitch. Excessive repeated loading can also cause fracture due to fatigue. From the design
model, the prototype spring seems to be at risk for this, but it is not certain maximum loading
occurs excessively in real use. Since a spring could still withstand a limited number of over
stressing load cycles, it must be established if this limit is reach in real use. This requires long term
field testing, and a suggested test procedure is outlined below.

• Before field testing, measure the free height of the spring and stiffness at the maximum
and minimum settings.
• Allow at least three test riders requiring close to the maximum stiffness set up one shock
and spring each and use them for at least 6 months without adjusting the stiffness.
• After the field test period is complete, remeasure the free height and stiffness at the
maximum and minimum settings.
• If possible, make use of a data logger during as much of the test period as possible to gather
data to estimate a representative load cycle for a shock spring.

In case of excessive over stressing, the tested springs would be permanently compressed, and the
free height reduced. If no significant reduction is measured for any of the springs after completed
testing, excessive recurrence of over stressing of the spring does likely not occur through regular
use. In that case, a higher wider adjustment range and increased limit for maximum shear stress
could possibly be used in following iterations of spring optimisation.

3.7 Analysis and model Verification


Verification of the spring model used was made by stress level comparison to results from FEM-
analysis using Solid Edge of the final design. Analytical analysis of a standard spring was also
performed to compare the level of over-stressing.

55
Numerical verification using FEA
A simplified model of the final spring was created with two open ends. This was done to capture
the behaviour of the active coils as this would be comparable to the analytical model used in the
design. Two models were used: one simulating minimum spring stiffness with 8,1 active coils and
one simulating maximum stiffness with 6,54 active coils.

An initial convergence study was performed to establish maximum allowable mesh size for mesh
independent results. This was done by fixing the bottom face of the spring wire and adding a
displacement of 65mm to the top face as illustrated in figure 35.

Figure 35. Constraints and loads in verification model


The mesh was refined, and maximum shear stress levels compared until a change in mesh sizing
caused less than 2% change in stress. Simulation results are presented in table. 12 and figure 36a)-
36d).

Table 12. Results from mesh convergence study


Mesh size Curvature 𝝉𝒎𝒂𝒙 𝚫𝝉
| 𝝉 𝒎𝒂𝒙 |
[mm] refinement ratio [MPa] 𝒎𝒂𝒙
[%]
10 0,5 1460 -
5 0,5 1130 22,6
2 0,5 1090 3,8
1 0,5 1090 <0,1

56
Figure 36. Results from convergence study with spring at minimum stiffness setting and maximum stroke
with a mesh size of a) 10mm, b) 5mm, c) 2mm, d) 1mm. For all results, a curvature refinement ratio of 0.5
was used.
As the stress levels showed mesh independence from a mesh size of 2mm this was the size chosen
for the rest of the analysis.

Each spring model were simulated with a displacement of 32,5mm and 65mm and the maximum
shear stress recorded. The actual deformed shape was also plotted to verify that no coil bind would
occur at maximum stroke. The results are shown in figure 37 and a comparison to analytical model
values in table 13.

57
Figure 37. Verification simulation results for a) minimum stiffness and 32.5mm stroke, b) minimum
stiffness and 65mm stroke, c) maximum stiffness and 32.5mm stroke, d) maximum stiffness and 65mm
stroke.

Table 13. Comparison of FEM stress to design model stress.


Active coils Stroke [mm] Max shear Max shear Design model
stress FEM stress design error [%]
[MPa] model [MPa]
8.1 32,5 551 523 5,1
8.1 65 1090 1047 3,9
6,54 32,5 672 648 3,6
6,54 65 1350 1297 3,9

A maximum deviation of 5% between the FEM and design model shear stress was found. The
design model is consistent in slightly underestimating the maximum shear stress in the spring wire.

58
However, the small deviation from the simulations results indicates that the design model is
accurate enough for the type of spring design in this project. The FE-model is likely to be more
accurate than the analytical models and therefore, the maximum allowed stiffness setting may have
to be lowered for the prototype spring.

Comparison with RockShox coil spring


In the unlikely event of both the FE-model and design model used being very inaccurate, a
reference maximum stress was established based on a standard spring from RockShox. The spring
has the following specifications:

• Rated stiffness: 105N/mm


• Rated maximum stroke: 55mm
• Inner diameter: 38,4mm
• Wire diameter: 8,8mm
• Number of active coils: 5
• Pitch: 23,5mm

Without knowing the steel type and shear modulus of the spring, it is impossible to determine the
accuracy of the rated stiffness. For comparison with the design models, it was therefore assumed
the rated stiffness was accurate within ±10% and the corrected maximum shear stress calculated.
The stress and stiffness were also calculated with the design model assuming a shear modulus of
79500MPa. The results are listed in table 14.
Table 14. List of calculated stresses for comparison between standard coil spring and adjustable
prototype spring
RockShox RockShox spring Final prototype
spring based assuming known spring
on rated shear modulus
stiffness
𝝉𝒎𝒂𝒙,𝒏𝒐𝒎𝒊𝒏𝒂𝒍 [MPa] 1310 1410 1300
𝝉𝒎𝒂𝒙,𝒖𝒑𝒑𝒆𝒓 [MPa] 1440 - -
𝝉𝒎𝒂𝒙,𝒍𝒐𝒘𝒆𝒓 [MPa] 1180 - -
𝒌𝒏𝒐𝒎𝒊𝒏𝒂𝒍 [N/mm] 105 113 87,2
𝒌𝒎𝒂𝒙 [N/mm] 115,5 - -
𝒌𝒎𝒊𝒏 [N/mm] 94,5 - -
𝑮𝒏𝒐𝒎𝒊𝒏𝒂𝒍 [MPa] 73600 79500 79500
𝑮𝒖𝒑𝒑𝒆𝒓 [MPa] 81000 - -
𝑮𝒍𝒐𝒘𝒆𝒓 [MPa] 66300 - -

From the nominal values, the RockShox spring is subject to the same level of maximum shear
stress as the adjustable spring. The stress in the adjustable spring lies just in middle of the estimated
range for the reference spring based on variability in its rated stiffness. Since the reference spring
is a production unit that has been on the market for many years, there are two conclusions that
could be drawn. Either, the use of a spring that is slightly over stressed at maximum shock travel
is possible on a mountain bike without severe issues. Or, the design model is inaccurate and too
conservative. The former is more likely due to the coherence with FE-modelling. In either case,
premature failure or severe issues should not be expected during field testing of the prototype
spring.

59
4 RESULTS
In this chapter, the results obtained from the concept development process are presented. These
are the results relevant to the research question. In addition, analysis of the results in context of
the theory presented in the frame of reference chapter is included.

4.1 Spring stiffness adjustment


Since manufacturing of prototypes and field testing could not be achieved within the timeline of
this thesis, the main results regarding spring stiffness adjustment are taken from the concept
development, evaluation, and design stages.

Most suitable spring type for an adjustable stiffness MTB shock


Based on the concept evaluation, stiffness adjustment could be achieved with several types of
mechanical springs, but the type best suited to current enduro mountain bikes is the helical coil
spring. It offers decent adjustment range, a desirable spring curve, excellent compatibility with
current bike frames and good environmental robustness. It can also be easily manufactured with
standard spring manufacturing technique and tooling. Apart from adjustability, the exclusive use
of coil springs in current mechanically sprung MTB shocks supports the evaluation of the coil
spring’s suitability for this purpose.

Achievable adjustment range


Adjustment of the helical coil spring is simply achieved by allowing one spring retainer to thread
along the spring. By altering the position of the retainer, the number of active coils can be
controlled and thus the spring stiffness. A linear relation between the number of active coils and
spring stiffness gives a linear spring adjustment in relation to the turns of the spring retainer
(adjuster).

As the stiffness is adjusted, the maximum stroke of the spring increases together with the
maximum shear stress in the wire for any given stroke. These factors limit the amount of
achievable adjustment. Material parameters affect the optimal design, so the results presented here
regarding achievable adjustment are based only on the specific properties of the material from
which the first prototypes are to be manufactured. This material is OTEVA 70 SC which is a spring
steel produced by Suzuki-Garphyttan with a shear modulus of 79500MPa and an ultimate tensile
stress of 1660-1760MPa. The achievable, or possibly achievable adjustment ranges based on
optimal spring design are presented in table 15.
Table 15. Results from spring design optimisation.
Min. Stiffness Max. Stiffness Adjustment Max. Stroke Max. % of
[N/mm] [N/mm] range [N/mm] @Max. predicted Ultimate
stiffness shear stress tensile
[mm] [MPa] strength
59,5 74,3 14,8 66,2 1150 69,3
70,4 87,2 16,8 66,7 1300 78,3
73,1 88,1 15 70,4 1310 78,9
64,6 83,2 18,6 66,5 1280 77,1

Each design allows a stroke of more than 66mm in their stiffest setting which makes them
compatible with a shock stroke of 65mm. Only one design gives a material stress below 70% of
the material UTS and thus clears the manufacturer’s design recommendation on maximum
allowable shear stress. This is true for a conservative estimate of the UTS based on the minimum
specified limit of 1660MPa.

60
If the actual strength is closer to the upper limit of 1760MPa, all the stiffness ranges specified in
table 15 are achievable with a max stress below 70% of the UTS. Based on this data and using a
conservative approach, only a spring with an adjustment range of 14,8N/mm (59,5-74,3N/mm) is
achievable with the specific material. This covers a rider weight range of approximately 59-75kg
with a maximum deviation of ±1% from the required maximum and minimum stiffness. Such a
weight range would accommodate for the middle 50% of the female Swedish population assumed
to approximate the middle 50% of female MTB riders. This does not come near the adjustment
range of current air shocks, but it is 70-110% greater than the range achieved by the only current
commercially available adjustable coil spring for mountain bikes. (The 70-110% depend on the
specific base rate of the Sprindex spring.)

With a less conservative estimate of the UTS, all the adjustment ranges in table 15 are achievable.
This means that an adjustable coil spring could theoretically be designed to accommodate riders
in the weight ranges 59-75kg, 74-89kg, 64-84kg, and 70-88kg. However, the variation in material
strength would likely cause all springs other than that for 59-75kg riders to be over stressed when
reaching maximum shock stroke at the stiffest setting.

61
5 DISCUSSION AND CONCLUSIONS
A discussion of the results and the conclusions that have been drawn during the Master of Science
thesis are presented in this chapter. The conclusions are based on the results and analysis with
the intention to answer the research question presented in the introduction chapter.

5.1 Discussion
The spring
The results show that it is theoretically possible to achieve a wide enough stiffness adjustment
range to accommodate a relatively wide range of rider weights by using an adapted coil spring.
Experimental verification must be performed however before it can be concluded to which extent
a spring can be theoretically over stressed while still being usable in real riding conditions.
Prototype testing is therefore of significant interest at this point. The maximum allowed stress limit
used to evaluate the coil spring was based on a design recommendation for springs in general. If
repeatedly stressed above this limit, a spring would break. Although an MTB shock spring would
reach its maximum design stress at full shock compression, it could be that a properly adjusted
shock does not reach full compression very often in real use. If that is the case, a spring could
potentially be designed with a higher maximum stress if the expected frequent maximum stress is
below the allowed limit. By extensive field testing with continuous shock position measurement,
an estimate of expected spring travel could be made. A typical representative compression cycle
for a shock spring could also be established. With such data, a maximum allowed spring stress
specific to MTB shock springs could be found which would help further optimise spring design
and find actual achievable adjustment ranges. In addition, a typical compression cycle could allow
for good estimates of spring fatigue life. This is an important factor in spring design which has
been omitted in this project due to lack of load cycle data.
A question that must be asked after the results have been compiled is: An adjustable coil spring
can likely be implemented for MTB use to suit a variety of riders, but should it? It offers great
advantage over conventional coil springs, and even an existing adjustable aftermarket coil spring
in terms of adjustment range. Having an adjustable spring that can accommodate riders in a weight
range of roughly 15kg is advantageous from a production point of view. Fewer springs must be
manufactured saving resources and costs associated with stock keeping. For the individual rider
however, the wide adjustment is of limited benefit compared to the drawbacks. An individual
requires a much narrower stiffness range for tuning to riding conditions. Once the correct spring
rate has been found, a rider most likely will not make use of more than an adjustment range of 5-
7N/mm, making about 50% of the available range superfluous. That is if the rider is in the middle
of the spring’s range. Riders closer to the limits of the range cannot use the adjustment equally in
each direction. The wide range becomes more beneficial for the individual though if the shock is
to be switched between several bikes with variations in leverage ratio and travel. Such a swap
could create a significant change in stiffness requirement which could be accommodated by a wide
spring adjustment range. From an overall perspective, an adjustable coil spring thus promotes a
more sustainable use of the shock where it can be reused on multiple bikes without requiring new
springs to be purchased for each.
The only significant drawback is the weight penalty that comes with the adjustability. The spring
to be manufactured for prototype testing, with an adjustment range of roughly 70-87N/mm was
estimated to weight 672g. Comparing this to an equivalent Sprindex spring, the weight penalty of
the increased adjustment is 226g. This is considered quite a lot in the MTB community, but not
excessive. For example, enduro riders often use tires with thicker casing to increase puncture
resistance. Such tires are roughly 150-200g heavier than most stock tires that come fitted when a

62
bike is purchased. A weight increase of 226g can thus be accepted if there is enough performance
increase.
If the achievable adjustment range would be enough to cover riders from 50-100kg for a similar
weight penalty, an adjustable coil spring would be ideal. It would completely remove the need for
production of multiple spring rates, reducing costs and use of resources. However, with the
achievable range established here, more detailed analysis of advantages and drawbacks should be
made before it could be definitively said if wide range adjustable coil springs should be used in
MTB. With a range of roughly 15N/mm, three base spring rates would likely cover for almost
every rider. From an environmental and economic perspective, this would be overwhelmingly
beneficial as only three, as opposed to around ten, springs would be required per shock model.
However, field testing and rider evaluation must be completed before it can be established whether
the weight penalty is justified by the benefits.
Shock integration
One issue that could potentially arise from manufacturing of the components in the shock and
spring is mismatch in the retainer-spring interface. If either the spring or the retainer cannot be
manufactured with accurate pitch and pitch diameter, misalignment and uneven spring support
occurs. The severity of this is difficult to predict and evaluation of the design must be performed
during prototype assembly and testing. If the prototype interface is not practically implementable,
a redesign is required.
Design model and optimisation
There are limitations to the models used to generate the final spring design in this project and these
must be highlighted. The accuracy of the analytical model used for finding spring stiffness and
shear stress has been verified using FEM and these models are not of concern in themselves.
However, simplifications and assumptions in the model used for defining the required adjustment
ranges may have introduced inaccuracies in the inputs to the spring optimisation. These
simplifications and assumptions have been justified by the limited time allowed for the project and
the model has been used to gain a representative estimate rather than exact data. Some
improvements could be proposed though to increase the accuracy of the estimates and the
confidence in the stiffness requirements.

A major assumption in the suspension model is that the rider mass is static and centred over the
bottom bracket. This is certainly not true during real MTB riding, but since the model aims to find
required spring stiffness based on static shock sag, this assumption still gives relatively accurate
results. If the same model should be made even more representative for mountain bikes, variation
in the static rider position could have been included. For example, if the normal static position for
a rider varies within ±10% around the bottom bracket, this could have been included to adjust the
minimum and maximum spring stiffness requirements to account for individual rider positions.

Another approach to make the model more representative in terms of riding position would have
been to base it on dynamic sag rather static. The static sag is typically used to find a good estimate
of the spring stiffness required by a rider. However, it is the dynamic handling of the bike that
ultimately dictates the spring stiffness requirement. Basing the model on dynamic sag could thus
account for riding styles and much better represent real use. Due to the vast variation of riding
styles, terrains and gradients involved in MTB, using the dynamic sag would drastically increase
the model complexity. First of all, extensive data acquisition would be required to measure shock
compression during riding. This would have to be done for several riders on different trail types
after having found their preferred stiffness settings. Thus, an ideal stiffness could be matched to
each combination of rider weight, riding style and terrain. Based on this an empirical model could
then be created that gives the ideal stiffness setting from rider weight, riding style and terrain as
input parameters. Frame geometry data could also be used in conjunction with the empirical model

63
to compensate the ideal spring stiffness for variations in relative bottom bracket position and
leverage ratio.

A significant simplification in the model was the use of average frame leverage ratio. Rarely any
bike today has a constant leverage ratio and most modern bikes have a falling leverage ratio
through the wheel travel. Thus, the initial leverage ratio is higher than the average and using the
average ratio would likely underestimate the required stiffness. Accounting for the actual frame
leverage curves would thus have made the model more accurate, although it would have required
linkage analysis of each individual bike. Also, it would not necessarily have made the results more
representative for all enduro bikes as all frames have different leverage curves and no standard
exists.
Final shock architecture
It is not very unexpected that the prototype spring concept is directly compatible with a
conventional shock layout. A more interesting shock design could potentially have been the result
if more design freedom had been given in terms of frame integration. The torsion spring concept
could have been beneficial if possible to integrate with a bike frame. One possible option would
be to place the torsion spring inside one of the frame tubes. This would both help seal the shock
from contaminations and reinforcements could easily be added for the frame to cope with the
torsional moment from the shock. Although ease of stiffness adjustment could be increased in this
way, it s likely that such a design would have significant drawbacks in terms of manufacturing
complexity.

5.2 Conclusions
An adjustable spring with a stiffness range suitable for the middle 50% of the Swedish female
population is achievable based on the results from this thesis. Also, other ranges suitable for the
middle 50% of the male population or the middle 50% of the combined population could
potentially be achievable after evaluation of field test results. This is based on a design that can
easily be implemented in a shock made for current MTB frames with standard shock mounts. This
then answers the research question of the thesis. Yes – it is possible to design a mechanical spring
with adjustable stiffness suitable for use on current mountain bikes by a wide variety of riders. It
is not possible to make an adjustable spring that suits all riders, but it is possible to achieve a wider
adjustment range than currently offered by available mechanical springs for MTB.
Determining the maximum allowable shear stress in a coil spring for use in MTB is difficult due
to the unpredictable and largely varying loading during real use. Spring design and stiffness
adjustment could thus be optimised further by establishing a representative load cycle for specific
MTB categories. This requires long-term field testing with data acquisition. The design of a coil
spring with wide adjustment range should therefore be an iterative process where field testing of
the first prototype designed in this thesis should establish the constraint parameters for the next
iteration of design optimisation.
Currently available spring steels do not offer the strength required to achieve an adjustable spring
suitable for all riders. To achieve the ultimate adjustable spring, other materials must therefore be
investigated for the application. Composites could potentially be a solution.

64
6 RECOMMENDATIONS AND FUTURE WORK
In this chapter, recommendations on more detailed solutions and future work in this field are
presented.

6.1 Recommendations
Due to limited project time, some detailed analysis was omitted from the design process. For the
sake of completeness, analysing end coil effects during spring compression and detailed damper
tuning could therefore be added. Also, optimisation of the damper body construction could be
performed to reduce overall weight. Detailed design of custom damper valve pistons could also
allow a reduction of the overall shock size.
For a better estimate of spring stiffness requirement, a model based on field test data should be
used. By gathering data about preferred spring stiffness, riding style and typical riding terrain from
a variety of riders would allow such factors to be accounted for when finding the ideal spring
stiffness. In addition, detailed linkage modelling would reduce inaccuracy in the model due to the
variably leverage ratio of modern enduro bikes.
If the design process is to be iterated, it should be preceded by data acquisition during filled testing
of the manufactured prototype, or even a standard coil spring. A representative load cycle and an
estimate of the number of cycles a shock spring is subject through its life should be established.
This would allow better estimate of maximum allowable spring stress in MTB shocks giving more
relevant optimisation constraints. It would also allow for a fatigue life estimate which is important
if the spring should be made available to customers.

6.2 Future work


Continuation and development of the work performed in this thesis project could be performed in
several areas. To improve the performance of the final spring design, other materials could be
investigated for the application.
The spring design optimisation should also be done for different shock size standards. Due to the
different ratios of shock length to stroke depending on the shock size, an adjustable spring may
well be better suited to another standard.
Adaptation of the concept to work in MTB suspension forks could also be beneficial. Coil springs
in suspension forks are becoming increasingly popular thanks to their linear spring curve and
sensitivity. The concept of an adjustable coil spring could therefore be even better suited to forks.
An analysis of shock mount designs in bike frames could also be conducted. Allowing the use of
torsion beam springs or different shock architecture thanks to a redesign in shock mounting could
potentially offer more adjustment.

65
7 REFERENCES

[1] Santa Cruz Bicycles, “Megatower,” 12 March 2021. [Online]. Available: https://www.santacruzbicycles.com/en-US/bikes/megatower.
[Accessed 29 July 2021].
[2] SRAM LLC., “Super Deluxe Coil Ultimate,” Rock Shox, 2021. [Online]. Available: https://www.sram.com/en/rockshox/models/rs-sdlc-
ult-a2. [Accessed 13 October 2021].
[3] FOX Factory, Inc., “DHX,” FOX Factory, Inc., 2021. [Online]. Available: https://www.ridefox.com/family.php?m=bike&family=dhx.
[Accessed 13 October 2021].
[4] B. Turman, “The Ultimate Guide to 'Metric' Shock Sizing and the RockShox Super Deluxe,” Vital MTB, 4 August 2016. [Online].
Available: https://www.vitalmtb.com/features/The-Ultimate-Guide-to-Metric-Shock-Sizing-and-the-RockShox-Super-Deluxe,1349.
[Accessed 15 October 2021].
[5] B. Morrison, “What's the deal with metric sizing,” amb Magazine, 17 January 2018. [Online]. Available:
https://www.ambmag.com.au/feature/whats-the-deal-with-metric-shock-sizing-478505/page0. [Accessed 13 Ocotber 2021].
[6] D. Goley, “EXT Storia Lok V3 Shock,” Blister, 22 September 2020. [Online]. Available: https://blisterreview.com/gear-reviews/ext-storia-
lok-v3-shock. [Accessed 13 October 2021].
[7] M. Alonso and A. Comas, “Modelling a twin tube cavitating shock absorber,” Proceedings of the Institution of Mechanical Engineers, pp.
1031-1040, 2006.
[8] E. Gelotte, “Development of software that can predict damper curves on shock absorbers,” Uppsala Universitet, Uppsala, 2013.
[9] K. S. RHOADES, “DEVELOPMENT AND EXPERIMENTAL VERIFICATION OF A PARAMETRIC MODEL OF AN
AUTOMOTIVE DAMPER,” Texas A&M University, College Station, 2006.
[10] P. Skačkauska, V. Žuraulis, V. Vadluga and S. Nagurn, “Development and verification of a shock absorber and its shim valve model based
on the force method principles,” Maintenance and Reliability, pp. 126-133, 2017.
[11] M. Kılıç, “Design and development of a mechanically adjustable linear torsion spring using cams,” Middle East Technical University,
Ankara, 2009.
[12] K. W. Hollander, T. G. Sugar and D. E. Herring, “Adjustable Robotic Tendon using a ‘Jack Spring’™,” in 2005 IEEE 9th International
Conference on Rehabilitation Robotics, Chicago, 2005.
[13] V. K. Doulatani and S. V. Chaitanya, “A Review on Technologies of Passive Suspension System with Variable Stiffness,” International
Journal of Engineering Research & Technology, pp. 917-921, 2014.
[14] Sprindex, “Sprindex - How it works / Progressive,” 18 February 2021. [Online]. Available: https://sprindex.com/pages/overview.
[15] Statistiska Centralbyrån, “Hälsa - Fler indikatorer 1980-2018,” 11 April 2019. [Online]. Available: https://scb.se/hitta-statistik/statistik-
efter-amne/levnadsforhallanden/levnadsforhallanden/undersokningarna-av-levnadsforhallanden-ulf-silc/pong/tabell-och-
diagram/halsa/halsa--fler-indikatorer/. [Accessed 13 March 2021].
[16] L. Uddhammar, Interviewee, Personal communication. [Interview]. 23 March 2021.
[17] Santa Cruz Bicycles, “Nomad,” 13 July 2021. [Online]. Available: https://www.santacruzbicycles.com/en-US/bikes/nomad. [Accessed 29
July 2021].
[18] Commencal Bikes, “Clash Signature 2021,” 29 June 2020. [Online]. Available: https://www.commencal-store.co.uk/clash-signature-
c2x31379299. [Accessed 29 July 2021].
[19] Commencal Bikes, “Meta AM 29 Signature Black/White 2021,” 19 July 2021. [Online]. Available: https://www.commencal-
store.co.uk/meta-am-29-signature-black-white-2021-c2x31636825. [Accessed 29 July 2021].
[20] YT-Industries, “Capra 29,” 12 March 2021. [Online]. Available: https://www.yt-industries.com/products/bikes/379/capra-pro-al-
29/preview. [Accessed 29 July 2021].
[21] Trek Bicycle Corporation, “Slash 9.7,” 2021. [Online]. Available: https://www.trekbikes.com/international/en_IN_TL/bikes/mountain-
bikes/trail-mountain-bikes/slash/slash-9-7/p/32994/?colorCode=orange_grey. [Accessed 29 July 2021].
[22] Trek Bicycle Corporation, “Remedy 8,” 2021. [Online]. Available: https://www.trekbikes.com/international/en_IN_TL/bikes/mountain-
bikes/trail-mountain-bikes/remedy/remedy-8/p/32883/?colorCode=grey. [Accessed 29 July 2021].
[23] R. G. Budynas, J. K. Nisbett and J. E. Shigley, “Chapter 10 Mechanical Springs,” in Shigley's Mechanical Engineering Design, 8th ed.,
London, McGraw-Hill, 2008, pp. 500-542.
[24] A. van Beek, “Chapter 14 - Machine design calculations reference guide,” in Advanced Engineering Design Lifetime Performance and
Reliability, Delft, Delft University of Technology, 2015, pp. 419-451.
[25] MakeItFrom, “MakeItFrom.com,” MakeItFrom, 30 May 2020. [Online]. Available: https://www.makeitfrom.com/material-
properties/ASTM-A227-Class-II-Spring-Steel. [Accessed 23 March 2021].
[26] D. POP, S. HARAGÂŞ and D. POPA, “HELICAL SPRING WITH VARIABLE WIRE DIAMETER,” Technical University of Cluj-
Napoca, Cluj-Napoca, 2017.
[27] B. Broberg, J. Carlsson, P.-O. Eriksson, G. Hedner, G. Härkegård, L.-E. Jarfall, H. Johansson, S. Kaiser, B. Larsson, S. E. Larsson, R.
Lundell, F. Nilsson, R. Nilsson, F. Niordson, F. K. Odqvist, A. Persson, G. Samnegård, B. Storåkers, B. Sundström, H. Strifors och B.
Åkesson, Handbok och formelsamling i Hållfasthestlära, Stockholm: Institutionen för hållfasthetslära KTH, 2014.
[28] M. Kazimer, “First Ride: The New EXT Era Fork is Very, Very Promising,” 30 July 2020. [Online]. Available:
https://www.pinkbike.com/news/first-ride-exts-new-era-fork-2021.html.
[29] DT Swiss, “COILPAR TEHCNOLOGY,” 13 July 2021. [Online]. Available: https://www.dtswiss.com/en/suspension/suspension-
technology/coilpair-technology.
[30] B. Dodson, P. C. Hammet and R. Klerx, “Desirability,” in Probabilistic design for optimization and robustness for engineers, Chichester,
John Wiley & Sons, Ltd, 2014, pp. 98-122.

66
[31] Motorex, “RACING FORK OIL 2,5W,” 2021. [Online]. Available:
https://motorex.com/Shared/Documents/TI/TIs%20PDF/MOTO%20LINE/Gabel-
Daempferoele/RACING%20FORK%20OIL/RACING%20FORK%20OIL%202,5W/RACING%20FORK%20OIL%202,5W_EN.pdf.
[Accessed 10 October 2021].
[32] Suzuki Garphyttan, “OTEVA 70 SC, OTEVA 70 SC PLUS,” 2018. [Online]. Available: https://www.suzuki-garphyttan.com/en/technical-
information/oteva/oteva-70-sc/. [Accessed 27 May 2021].
[33] Siemens, Solid Edge 2020, Siemens, 2020.
[34] SRAM LLC., “Super Deluxe Ultimate,” 2021. [Online]. Available: https://www.sram.com/en/rockshox/models/rs-sdlx-ult-b2. [Accessed
13 October 2021].
[35] FOX Factory, Inc., “FloatX,” 2021. [Online]. Available: https://www.ridefox.com/family.php?m=bike&family=floatx. [Accessed 13
Ocotber 2021].
[36] FOX Factory, Inc., “2021-2022 FLOAT X2 User Specifications,” FOX Factory, Inc., 2021. [Online]. Available:
https://www.ridefox.com/fox17/help.php?m=bike&id=1094. [Accessed 19 October 2021].

67

You might also like