77
Numerical Modelling of a Radial Inflow Turbine with and without Nozzle Ring at Design and Off-Design Conditions Filippo Valentini Master of Science Thesis EGI_2016-094 MSC EKV1169 KTH School of Industrial Engineering and Management Machine Design SE-100 44 STOCKHOLM

Numerical Modelling of a Radial Inflow Turbine with and

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Numerical Modelling of a Radial Inflow Turbine with and

Numerical Modelling of a Radial Inflow Turbine with and without Nozzle Ring at Design and Off-Design Conditions

Filippo Valentini

Master of Science Thesis EGI_2016-094 MSC EKV1169

KTH School of Industrial Engineering and Management

Machine Design

SE-100 44 STOCKHOLM

Page 2: Numerical Modelling of a Radial Inflow Turbine with and

Master of Science Thesis

EGI_2016-094 MSC EKV1169

Numerical Modelling of a Radial Inflow Turbine with and without Nozzle Ring at Design and Off-

Design Conditions

Filippo Valentini

Approved

Examiner

Paul Petrie-Repar

Supervisor

Jens Fridh

Commissioner

Contact person

Abstract

The design of a radial turbine working at peak efficiency over a wide range of operating

conditions is nowadays an active topic of research, as this constitutes a target feature for

applications on turbochargers. To this purpose many solutions have been suggested, including

the use of devices for better flow guidance, namely the nozzle ring, which are reported to boost

the performance of a radial turbine at both design and off-design points. However the majority of

performance evaluations available in literature are based on one-dimensional meanline analysis,

hence loss terms related to the three-dimensional nature of real flows inside a radial turbine are

either approximated through empirical relations or simply neglected.

In this thesis a three-dimensional approach to the design of a radial turbine is implemented, and

two configurations, with and without fixed nozzle ring, are generated. The turbine is designed for

a turbocharging system of a typical six-cylinder diesel truck engine, of which exhaust gas

thermodynamic properties are known. The models are studied by means of a CFD commercial

software, and their performance at steady design and off-design conditions are compared.

Results show that, at design point, the addition of a static nozzle ring leads to non negligible

increments, with respect to the vaneless case, of both efficiency and power output: such

increments are estimated in +1.5% and +3.5% respectively, despite these data should be

compared with the uncertainty of the numerical model. On the other hand both turbine

configurations are found to be very sensitive to variations of pressure and temperature of the

incoming fluid, hence off-design performances are dependent on the particular off-design point

considered and a “best” configuration within all the combustion cycle does not exist.

Page 3: Numerical Modelling of a Radial Inflow Turbine with and

ACKNOWLEDGEMENTS

I would like to express my gratitude to my supervisor, Jens Fridh, and my examiner, Paul Petrie-

Repar, for their availability and support, and without whom this work would have never been

realised.

Sincere gratitude to my Italian examiner Alessandro Talamelli, who followed the project of Dual

Degree between UniBo and KTH since its inception: it is also thanks to him that I ended up at

KTH, a circumstance which I will never regret.

I would like to thank my parents, who believed in me and allowed me to complete my academic

iter, providing full moral and material support.

A special mention goes to my Italian friends: Pietro, who was always ready to support me with a

Skype call and welcome me back during my short visits in Italy, and Simone, who, no matter

where he is, always finds a way to be present in the most crucial moments.

Thank you Danilo, you proved to be not only a respectful roommate but also a real valuable

person from all sides. And many thanks to all friends that I met in Stockholm, who turned my

stay abroad into an exciting and enjoyable experience, because “there’s no point in living if you

can’t feel alive”.

Page 4: Numerical Modelling of a Radial Inflow Turbine with and

iii

TABLE OF CONTENTS

List of Figures.................................................................................................................................v

List of Tables.................................................................................................................................vii

Nomenclature...............................................................................................................................viii

1. Introduction...............................................................................................................................1

1.1. The Radial Turbine.............................................................................................................2

1.2. Sources of Losses...............................................................................................................5

1.3. The Design Process............................................................................................................7

1.4. Design and Analysis: State of the Art...............................................................................8

1.4.1. Inverse Problem: the Volute...................................................................................8

1.4.2. Inverse Problem: the Rotor.....................................................................................9

1.4.3. Inverse Problem: the Remaining Components......................................................10

1.4.4. Direct Problem: Performance Analysis.................................................................10

2. Motivation and Objective........................................................................................................11

2.1. Motivation........................................................................................................................11

2.2. Objective..........................................................................................................................11

3. Methodology and Tools..........................................................................................................12

3.1. Methodology....................................................................................................................12

3.2. Tools-Software.................................................................................................................13

4. Limitations...............................................................................................................................15

5. Design of Components............................................................................................................16

5.1. Design of the Volute........................................................................................................16

5.1.1. Theoretical Procedure...........................................................................................17

5.1.2. Implementation of the Theoretical Procedure.......................................................19

5.1.3. Implementation Under Additional Constraints.....................................................21

5.2. Design of the Nozzle Ring...............................................................................................23

5.3. Design of the Rotor..........................................................................................................25

5.3.1. Preliminary Design................................................................................................25

5.3.2. The Bezier Curve...................................................................................................27

5.3.3. Implementation of the Design Strategy.................................................................28

5.3.4. Supplementary Issues on 3D Design of the Rotor.................................................31

5.4. Design of the Diffuser......................................................................................................33

6. Mesh Generation......................................................................................................................35

6.1. Choice of the Grid............................................................................................................35

Page 5: Numerical Modelling of a Radial Inflow Turbine with and

iv

6.2. Meshing the Boundary Layer...........................................................................................36

6.3. Quality of the Mesh..........................................................................................................38

6.3.1. Skewness................................................................................................................38

6.3.2. Orthogonal Quality................................................................................................39

6.3.3. Jacobian Ratio........................................................................................................39

6.4. Meshing of Components...................................................................................................40

7. CFX® Setup..............................................................................................................................46

7.1. Mathematical Model for Turbulence................................................................................46

7.2. Near-Wall Treatment........................................................................................................47

7.3. Boundary Conditions and Interfaces...............................................................................47

7.4. Choice of Off-Design Points............................................................................................48

8. Results.....................................................................................................................................50

8.1. Design Point.....................................................................................................................50

8.2. Off-Design Points............................................................................................................54

8.2.1. Off-Design Point 1................................................................................................55

8.2.2. Off-Design Point 4................................................................................................56

8.2.3. Off-Design Point 5................................................................................................57

8.2.4. Off-Design Point 7................................................................................................58

9. Discussion...............................................................................................................................59

10. Conclusions and Future Works...............................................................................................61

11. Bibliography............................................................................................................................62

Appendix 1: Absolute Angle at Rotor Inlet..................................................................................65

Page 6: Numerical Modelling of a Radial Inflow Turbine with and

v

LIST OF FIGURES

NUM. TITLE PAG.

Figure 1 Thermodynamics of a radial turbine (Nguyen-Schäfer, [16], adapted) 1

Figure 2 Scheme of a 90° IFR turbine. Left: frontal view. Right: side view (Ventura,

[13])

2

Figure 3 Flow velocity triangles within a radial turbine (Dixon, [6]) 3

Figure 4 Thermodynamic diagram of the process through a 90° IFR turbine (Dixon,

[6])

4

Figure 5 Nominal design configuration (Saravanamuttoo, [19], adapted) 5

Figure 6 Behaviour of loss terms as function of incidence angle (Yahya, [33]) 6

Figure 7 Secondary flow in a blade passage (Yahya, [33]) 7

Figure 8 Full process iter in turbine design (Khader, ,[9]) 8

Figure 9 Schematic diagram of a vaneless volute casing (Whitfield, [31]) 16

Figure 10 Left – velocity profile across the centre-line of the volute section. Right –

variation of centroid radius at two subsequent azimuth positions (Whitfield,

[31], adapted)

18

Figure 11 Theoretical distribution of centroid radius, cross section area and flow

angle with azimuth location

20

Figure 12 Distribution of centroid radius, cross section area and flow angle with

azimuth location. Comparison between theoretical and implemented

solution, vaneless case

21

Figure 13 Distribution of centroid radius, cross section area and flow angle with

azimuth location. Comparison between theoretical and implemented

solution, vaned case

22

Figure 14 3D geometrical model of the volute casing. Left: vaneless. Right: vaned (no

nozzle)

23

Figure 15 Nozzle vane geometry definition (Rajoo & Martinez-Botas, [18]) 24

Figure 16 Nozzle ring. Left: sketch in the frontal plane. Right: shape of the blade 24

Figure 17 3D geometrical model of the nozzle ring 25

Figure 18 Sketch of the velocity triangles at rotor inlet and outlet (Saravanamuttoo,

[19], adapted)

26

Figure 19 Rotor views. Left: r-z (or meridional) plane. Right: θ-z (or blade to blade)

plane

27

Figure 20 Bezier curve of degree 3. Left: basis of vector space. Right: control points

(Floater, [7], adapted)

28

Figure 21 Rotor: θ-distribution (top), β-distribution (middle), thickness distribution

(down)

29

Figure 22 Rotor: distribution of wrap angle (top left), flow angle (top right) and

thickness (bottom) in the meridional plane

30

Figure 23 Rotor: variation from inlet to outlet of channel cross-section area (left) and

lean angle (right)

31

Figure 24 3D geometrical model of the rotor 31

Figure 25 Geometrical definition of the problem 32

Figure 26 Pressure distribution in a crosswise section. Effect of streamwise pressure

gradient (left), effect of blade-to-blade pressure gradient (middle),

ensemble (right) (Van den Braembussche, [28])

33

Figure 27 Conical diffuser. Left: 2D sketch. Right: lines of appreciable stall for given

geometrical configuration (Blevins, [4], adapted)

34

Figure 28 3D geometrical model of the diffuser 34

Page 7: Numerical Modelling of a Radial Inflow Turbine with and

vi

Figure 29 Elements of a 3D mesh - tetrahedron, hexahedron, prism, pyramid 35

Figure 30 Velocity profile in a turbulent boundary layer (Bakker, [3]) 36

Figure 31 Non-dimensional velocity as function of 𝑦+ in the inner region (Kundu,

[11])

37

Figure 32 Stretching of a quadrilateral element. Nominal shape (left), deformed

shape (right)

38

Figure 33 Orthogonal quality on a 2D quadrilateral cell 39

Figure 34 Mapping of an hexahedral element (Bucki, [5]) 39

Figure 35 Example of setup of 𝐴𝑁𝑆𝑌𝑆®Meshing (volute) 40

Figure 36 Mesh of the volute (section) 42

Figure 37 Mesh of the diffuser (ensemble) 42

Figure 38 Mesh of the nozzle ring (one blade) 43

Figure 39 Topology for the rotor blade. The blade (blue) is surrounded by meshing

blocks

43

Figure 40 Mesh of the rotor (portion) 44

Figure 41 Mesh statistics for the rotor 45

Figure 42 Illustration of interfaces and boundary conditions for vaned configuration 48

Figure 43 Choice of representative off-design points (Mora, [1], adapted)

49

Figure 44 Spanwise distribution of 𝛽2, comparison at design point 51

Figure 45 Spanwise distribution of 𝛼3, comparison at design point 51

Figure 46 Rotor blade loading comparison at 10% span (top, left), 50% span (top,

right), 90% span (bottom), design point

52

Figure 47 Static entropy around the blade at 90% span. Left: vaneless. Right: vaned 52

Figure 48 Mach distribution around the blade at 10% span. Top: vaneless.

Bottom:vaned

53

Figure 49 Mach distribution around the blade at 50% span. Left: vaneless. Right:

vaned

53

Figure 50 Mach distribution around the blade at 90% span. Left: vaneless. Right:

vaned

53

Figure 51 Velocity contour around nozzle ring blade, 50% span, design point 54

Figure 52 Spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 1 55

Figure 53 Velocity distribution around the blade at 50% span. Left: vaneless. Right:

vaned

55

Figure 54 Velocity distribution around the blade at 90% span. Left: vaneless. Right:

vaned

55

Figure 55 Spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 4 56

Figure 56 Static entropy in meridional lane, circumferential average, off-design point

4. Left: vaneless. Right: vaned

56

Figure 57 Velocity distribution around the blade (left) and blade loading (right) at

90% span, off-design point 4

56

Figure 58 Spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 5 57

Figure 59 Static entropy in meridional plane, circumferential average, off-design

point 5. Left: vaneless. Right: vaned

57

Figure 60 Velocity contour around nozzle ring blade, 50% span, off-design point 5 57

Figure 61 Spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 7 58

Figure 62 Rotor blade loading, comparison at 10% span (top, left), 50% span (top,

right), 90% span (bottom), off-design point 7

58

Figure 63 spanwise distribution of 𝛼2. From top to bottom: design point, off-design 1,

off-design 4, off-design 5, off-design 7

66

Page 8: Numerical Modelling of a Radial Inflow Turbine with and

vii

LIST OF TABLES

NUMBER TITLE PAGE

Table 1 Meanline design parameters for nozzle ring 19

Table 2 Meanline design parameters for vaneless and vaned volutes 20

Table 3 Relative angle at rotor outlet under nominal design condition 26

Table 4 Meanline design parameters for the rotor 28

Table 5 Meanline design parameters for the diffuser 33

Table 6 Estimation of first layer thickness for turbine components 41

Table 7 Mesh statistics for turbine components 41

Table 8 Thermodynamic properties of the studied off-design points 49

Table 9 Performance comparison, design point 50

Table 10 Comparison between mean velocity triangles at rotor inlet (top) and at

rotor outlet (bottom), design point

50

Table 11 Comparison of performance (top) and mean flow angles (bottom) at off-

design points

54

Page 9: Numerical Modelling of a Radial Inflow Turbine with and

viii

NOMENCLATURE

ABBREVIATIONS

ATM Automated Topology and Meshing

CAD Computer Aided Design

CFD Computational Fluid Dynamics

DNS Direct Numerical Simlation

FEA Finite Element Analysis

FEM Finite Element Method

IFR Inflow Radial

JR Jacobian Ratio

LE Leading Edge

PDE Partial Differential Equation

RANS Reynolds Averaged Navier-Stokes

RPM Revolutions per Minute

OQ Orthogonal Quality

SBP Single Blade Passage

SST Shear Stress Transport

TE Trailing Edge

T-T Total-to-Total

1-2-3D One-Two-Three Dimensional

GREEK SYMBOLS

α Absolute flow angle

Constant (k- ω model)

β Relative flow angle

Blade angle (rotor)

Constant (k- ω model)

γ Specific heat ratio

δ Boundary layer thickness

ε Dissipation of turbulent kinetic energy

η Efficiency

θ Wrap angle (rotor)

Non-dimensional mass flow rate (volute)

Deformation angle (mesh statistics)

λ Angle (general notation)

μ Dynamic viscosity

μ𝑡 Turbulent viscosity

ν Kinematic viscosity

ξ General coordinate

ρ Density

ζ Constant (k- ω model)

η Shear stress

Φ, θ Azimuth angle (volute)

ψ Diffusion angle (diffuser)

Ω, ω Angular velocity

Dissipation rate of turbulent kinetic energy (RANS model)

ROMAN SYMBOLS

Page 10: Numerical Modelling of a Radial Inflow Turbine with and

ix

a Speed of sound

A Area

b Passage width (volute)

Vane axial chord (nozzle ring)

c Absolute flow velocity

Vane true chord (nozzle ring)

𝐶𝑓 Skin friction coefficient

g Gravitational acceleration

h Enthalpy

Blade height (volute)

i,j,k Indexes (general notation)

k Constant (volute)

Turbulent kinetic energy (RANS model)

𝑘𝑎 Axial length (rotor)

l Axial length (diffuser)

L Work exchange

m Meridional coordinate

Vortex exponent (volute)

Mass (general notation)

𝑚 Mass flow

M Mach number

Moment of external forces (general notation)

n Rotational speed

Coordinate normal to streamline (rotor)

o Nozzle throat

P, p Pressure

Constant (volute)

q Heat exchange

R, r Radius

𝑅𝑒 Reynolds number

s Entropy

Nozzle pitch (nozzle ring)

S Angular momentum ratio (volute)

T Temperature

u Velocity (general notation)

U Blade speed

Velocity (general notation)

𝑢∗ Friction velocity

w Relative flow velocity

W Work exchange

X, x General coordinate

y General coordinate

𝑦+ Non-dimensional coordinate normal to a wall

𝑍𝑣 Number of blades (nozzle ring)

𝑍𝑏 Number of blades (rotor)

SUBSCRIPTS (*)

0 Stagnation property

1 At volute inlet

2 At volute outlet, At rotor inlet

Page 11: Numerical Modelling of a Radial Inflow Turbine with and

x

3 At rotor outlet, At diffuser inlet

4 At diffuser outlet

a Relative to axial coordinate

b Relative to the blade

e Relative to ideal case (mesh statistics)

h Relative to blade hub (rotor)

i,j,k Relative to indexes i,j,k

m Relative to meridional coordinate

n Relative to the coordinate n

nr At nozzle ring

r Relative to radial coordinate

R Recirculating at volute tongue (volute)

ref Relative to a reference quantity

s Along isentropic transformation, Static

t Stagnation property

Relative to blade tip (rotor)

Relative to the tangential coordinate (rotor)

ts Total-to-static

tt Total-to-total

x Relative to the coordinate x

y Relative to the coordinate y

θ Relative to circumferential coordinate

(*) Some subscripts used in the notation are self-explanatory (i.e. inlet, wall, max…) and are not listed here

Page 12: Numerical Modelling of a Radial Inflow Turbine with and

1

1 – INTRODUCTION

In an internal combustion engine a turbocharger is a system composed by a radial turbine and a

centrifugal compressor mounted on a common shaft. The turbine converts part of the enthalpy of

the exhaust gases into kinetic energy delivered to the shaft. The shaft drives the compressor

which increases the air pressure sent to the combustion chamber at each piston cycle, allowing

an increase in power output: for this reason turbocharging is often considered in applications

where power demand is a priority, such as in race car, heavy duty vehicle and marine engines.

Turbocharging is also beneficial from the point of view of overall engine efficiency. As noted by

Mora [14], more power available per unit cycle leads to smaller engines, with consequent

decrease of mechanical losses, and the higher density of air at the combustion chamber improves

volumetric efficiency. Moreover the performance of a supercharged engine is only minimally

affected by variations in ambient pressure, since pressure is a design parameter of the

turbocharger itself at steady conditions: consequently the engine operates at more constant

regime.

The absolute performance of a turbocharger is limited by the energy content of the working

fluid, defined by its total enthalpy. As shown in Fig.1, the work that the turbine can extract is the

difference between the total enthalpy at rotor inlet and the total enthalpy at rotor outlet, i.e.

without taking into account the contribution of flow speed. Whatever velocity component the

flow still has at outlet, infact, is “wasted” from a thermodynamic point of view; in the ideal case

the turbine should be able to expand the flow so to reach outlet pressure at zero velocity, and

without introducing losses (i.e. isentropically).

Figure 1: thermodynamics of a radial turbine (Nguyen-Schäfer, [16], adapted)

The ratio between the actual work and the total available energy defines the efficiency of the

turbine, which is a key driver in the design of the turbomachinery.

In this introductory part a description of the radial turbine is given: focus is placed on the

thermodynamic process underwent by the flow through all components of the turbine and on

Page 13: Numerical Modelling of a Radial Inflow Turbine with and

2

physical aspects of power generation. Then the main dissipative phenomena arising in a real-case

turbomachinery are presented, since their knowledge is crucial for design purpose. The

subsequent section illustrates the steps of the design process, which also the present work is

based on. The introduction is concluded by a literature review highlighting the results obtained

so far in the design of radial turbines.

1.1 The radial turbine

For turbocharger applications a 90° IFR turbine is generally employed, as it guarantees

compactness, good efficiency within a wide range of operating conditions and high structural

strength. A sketch of such a turbine is shown in Fig.2.

Inside the volute the flow is accelerated and at the outlet it is delivered uniformly with a desired

outflow angle. At this stage the flow can enter directly into the rotor or pass through a nozzle

ring (stator): the former configuration is called vaneless, the latter vaned. The nozzle ring, when

present, aims at guiding the flow into the rotor inlet, and since the passage between blades is

convergent it further increases its speed. Inside the rotor (also called impeller) the flow exerts an

aerodynamic force on the blades, thus transferring part of its energy to the rotating device: the

goal of the impeller is to extract the highest possible amount of work from the working fluid.

The flow exits the impeller with a mainly axial velocity component and passes through a

diffuser, where it is slowed down and part of its static pressure is recovered before being

discharged.

Figure 2: scheme of a 90° IFR turbine. Left: frontal view. Right: side view (Ventura, [13])

The process can be described from a thermodynamic point of view, which is useful to highlight

the physical quantities affecting the performance of the turbine and to deduce preliminary

considerations about its design.

Most turbomachinery flow processes are adiabatic (Dixon, [6]), and also gravitational effects are

negligible; the first equation of thermodynamics for an open system (Negri di Montenegro, [15])

may then be re-written in terms of total enthalpy 𝑕0 as shown in eqn.(1.1)

Page 14: Numerical Modelling of a Radial Inflow Turbine with and

3

𝑑𝑕 + 𝑐𝑑𝑐 + 𝑔𝑑𝑧 = 𝑑𝑞 − 𝑑𝐿 𝑑𝑞=0; 𝑑𝑧=0; −𝑑𝐿=𝑑𝑊 𝑑𝑕 + 𝑐𝑑𝑐

𝑑 𝑕+𝑐2

2

= 𝑑𝑊 𝑕0≝𝑕+

𝑐2

2 𝑑𝑕0 = 𝑑𝑊 (1.1)

The work per unit mass exerted by the flow on the turbine shaft is linked to the rate of change of

angular momentum that the flow itself undergoes inside the rotor, and may be expressed by the

so called “Euler turbine equation”, here in differential form (for the exact derivation see Mora

[14])

𝑑𝑊 = 𝑑 𝑈𝑐𝜃 (1.2)

The combination of eqn.(1.1) and eqn.(1.2) yields to eqn.(1.3):

𝑑(𝑕0 − 𝑈𝑐𝜃) 𝑟𝑜𝑡𝑕𝑎𝑙𝑝𝑦

= 0 (1.3)

which states that in the thermodynamic process through the impeller the rotational total enthalpy

𝐼 ≝ 𝑕0 − 𝑈𝑐𝜃 (also called rothalpy) is constant. The same expression is re-arranged according

to the velocity triangle in Fig.3, which relates the velocity vectors in the fixed and the rotating

frame of reference:

𝐼 = 𝑕 +1

2𝑐2 −𝑈𝑐𝜃 = 𝑕 +

1

2 𝑤2 + 𝑈2 + 2𝑈𝑤𝜃 − 𝑈 𝑤𝜃 + 𝑈 = 𝑕 +

1

2𝑤2 −

1

2𝑈2 (1.4)

Figure 3: flow velocity triangles within a radial turbine (Dixon, [6])

Through eqn.(1.4) it is possible to express the variation of static enthalpy between inlet and

outlet of the rotor (denoted by indexes 2 and 3, Fig.4), thus from the combination of eqn.(1.1)

with eqn.(1.4) the total specific work on the turbine is:

∆𝑊2−3 = 𝑕 +𝑐2

2

2−3=

1

2[ 𝑈2

2 − 𝑈32 − 𝑤2

2 − 𝑤32 + 𝑐2

2 − 𝑐32 ] (1.5)

Eqn.(1.5) leads to the following considerations:

Page 15: Numerical Modelling of a Radial Inflow Turbine with and

4

a) In the case of an IFR turbine the term 1

2 𝑈2

2 − 𝑈32 is always positive because 𝑈2 > 𝑈3; the

same does not hold for an axial turbine, where inlet and outlet radiuses are equal. As

direct consequence, the work per unit mass that can be extracted with a single turbine

stage is higher in the former case, and so the efficiency, which justifies the exclusive

employment of radial turbines in turbochargers.

b) A positive contribution to the specific work is obtained when 𝑤3 > 𝑤2. This is achieved

if the channels between blades are convergent.

Eqn.(1.4) shows that the higher is 𝑤3 the higher is the static enthalpy jump 𝑕2 − 𝑕3 across the rotor (and consequently the specific work) but the lower is the static pressure 𝑝3 at rotor outlet (see Fig.4). However 𝑝3 cannot be lower than the atmospheric pressure,

otherwise the discharge would not be possible. The diffuser is used to allow 𝑝3 < 𝑝𝑎𝑡𝑚 (with a benefit for efficiency) and to recover part of the static pressure afterwards so that

at turbine outlet the condition 𝑝4 ≥ 𝑝𝑎𝑡𝑚 is fulfilled.

Moreover “accelerating the relative velocity through the rotor is a most useful aim of the

designer as this is conducive to achieving a low loss flow” (Dixon, [6]).

c) Since the specific work is proportional to the quantity 1

2(𝑐2

2 − 𝑐32) the absolute velocity

should be large at impeller inlet, which is achieved by means of the volute.

The volute (as well as the diffuser) is a static component (𝑑𝑊 = 0) and from eqn. (1.1)

the total enthalpy is conserved (see also Fig.4) which implies that the higher is the drop in

static enthalpy 𝑕1 − 𝑕2 the higher is the absolute velocity seen by the rotor at its inlet.

Figure 4: thermodynamic diagram of the process through a 90° IFR turbine (Dixon, [6])

With reference to point c) above, a deeper analysis is needed. Eqn.(1.2) can be evaluated

between points 2 and 3, leading to eqn.(1.6)

∆𝑊2−3 = 𝑈2𝑐𝜃2 − 𝑈3𝑐𝜃3 (1.6)

which shows that:

Page 16: Numerical Modelling of a Radial Inflow Turbine with and

5

a high value of 𝑐2 is not the only requirement because what matters is the projection of 𝑐2

along the circumferential direction, 𝑐𝜃2. The ideal condition would be 𝑐2 ≡ 𝑐𝜃2 but this is

not physical since there would be no inflow at rotor inlet: the absolute flow angle is

chosen such that the relative velocity has only radial component.

A good volute design must therefore take into account the orientation of the absolute

velocity at rotor inlet, which may be accomplished by using vaned stators.

the specific work is increased if the absolute velocity of the flow at rotor outlet is axial,

i.e. 𝑐𝜃3 = 0.

The two conditions mentioned above constitute the so called nominal design (Fig.5).

Figure 5: nominal design configuration (Saravanamuttoo, [19], adapted)

1.2 Sources of losses

Nominal design is a theoretical condition which sets an upper bound to the maximum specific

work that a radial turbine can extract.

In real cases, however, this limit is never reached because dissipative phenomena arise from the

interaction between flow and solid surfaces of the turbomachinery: these phenomena (take Fig.1

as reference) increase the level of entropy and hence lower the jump in total enthalpy across the

turbine, which is linked to the amount of work by eqn.(1.1).

Several loss models have been developed and are now available in literature (Ventura, [13])

since careful evaluation of losses is crucial for performance estimation: however most of them

are based on empirical relations. In this section a qualitative description of the main sources of

losses in radial turbines is presented, and observations are made on how to limit them.

Channel losses

Dissipative phenomenon due to skin friction. Skin friction depends on wet surface

(portion of the surface in contact with the flow), surface roughness and flow speed: a

way to limit channel losses is then to make compact turbines with surfaces as smooth as

possible (limiting the flow speed would infact be detrimental for the expansion process

inside the rotor)

Page 17: Numerical Modelling of a Radial Inflow Turbine with and

6

Incidence losses

Here are considered, without further distinctions, all losses which arise from the non-

ideal orientation of the velocity vector with respect to the rotor blades (and blades of the

nozzle ring, when present). For a 90° IFR turbine incidence losses originate when the

flow does not enter the rotor radially or leave it axially, as stated in the nominal design.

Incidence losses (also called shock losses, even if they are not related to compressibility

effects) grow almost parabolically with incidence angle (Dixon, [6]): the specific

behaviour depends on the loss model adopted but more in general Fig.6 shows that this

term contributes heavily to overall losses at off-design incidence angles.

Figure 6: behaviour of loss terms as function of incidence angle (Yahya, [33])

A well known solution to reduce incidence losses is to achieve better flow guidance at the

exit of the volute by means of a nozzle ring.

Secondary losses

Secondary flow generally denotes a portion of fluid which does not follow the streamlines

of the main flow.

As an example let us consider a channel between rotor blades (Fig.7). Suction and

pressure sides of adjacent blades create a pressure gradient orthogonal to the streamlines:

this induces a crossflow which develops a boundary layer in the local cross section planes

of the channel. A direct consequence is the onset of shear stresses because of the velocity

gradient (recall 𝜏 = 𝜇 𝜕𝑈

𝜕𝑦 𝑤𝑎𝑙𝑙

); moreover the vorticity injected in the flow creates

eddies which dissipate kinetic energy through turbulent mechanism (energy cascade).

Another region in which the presence of secondary flows is not negligible is in the

neighbourhood of the volute tongue, i.e. the connection between the inlet and the end of

the volute casing, where the flow has a lower pressure than at inlet because it was

accelerated inside the volute itself. The resultant pressure difference drives part of the

flow from the inlet around the tongue, creating a zone of recirculation thus causing

losses.

Page 18: Numerical Modelling of a Radial Inflow Turbine with and

7

Figure 7: secondary flow in a blade passage (Yahya, [33])

Tip clearance losses

Tip clearance is a gap between blade shroud and casing which is designed to allow the

expansion of the rotor under thermal effect and centrifugal force. Across this gap leakage

occurs between the pressure and the suction side of the blade, driven by the pressure

gradient: the behaviour is the same of a wing of finite span.

This secondary flow through the clearance does not contribute to the work done on the

rotor, causing a reduction of the work output compared to the condition of designed mass

flow. Moreover tip clearance flow is responsible for the formation of vortices which may

also interact with other secondary flows (Siggeirsson et al., [22]).

Tip clearance losses depend on the relative size of the gap compared to the blade height.

1.3 The design process

In section 1.1 an ideal design condition was derived from studying the thermodynamics of the

expansion process inside the radial turbine. As this approach does not require any knowledge

about the internal flow pattern in terms of velocity and pressure fields, but only the average

physical quantities at specific sections of the turbomachinery, it is a common tool at early stages

of the design process: such approach is called meanline design.

Meanline analysis consists of treating the working fluid as a one dimensional flow at the mean

radius of the turbine while the flow parameters are assumed as reasonable average values across

the full span (Wei, [29]).

The flow in a turbine is, however, fully three-dimensional. As it was shown in the previous

section, some dissipative phenomena, like losses due to secondary flows, have an inherent three-

dimensional nature and meanline design would not be able to detect them. The next step is then

to develop a complete geometrical model which should yield to an expected behavior of the

internal flow (taking into account secondary flows as well, if their modelization is available) and

which, on average, is expected to have a performance close to the one forecasted with the

meanline design. Due to high complexity of the internal flow and the amount of constraints to be

fulfilled, well established procedures for this phase of the design process are difficult to develop,

and few results are reported in literature: a key role is still played by experience and personal

know-how of the designer.

Page 19: Numerical Modelling of a Radial Inflow Turbine with and

8

Finally an accurate performance estimation is done on the 3D geometrical configuration by

simulating the internal flow with CFD techniques. A feasibility check is needed as well in order

to assess structural robustness under operating loads: FEA is commonly employed in this case.

Figure 8: full process iter in turbine design (Khader, ,[9])

“Turbine design is an iterative process” (Khader, [9]). Fig.8 shows the correlation between

different steps: in particular the initial 3D geometry is modified according to the outcomes of the

analysis step. The latter are continuously compared with the results from the meanline design

(which is an idealized design, thus represents the target) and the final configuration is reached

when sufficient convergence exists between the two steps.

1.4 Design and analysis: state of the art

“There are two problems of interest to designers of turbomachinery” (Yang, [34]). In the so

called “inverse” (or design) problem the overall geometry is defined, based on assumptions

about the flow pattern and other existing specifications, to yield desired performance

characteristics, while in the “direct” (or analysis) problem the performance of a given

geometrical configuration is assessed.

The direct problem is commonly tackled by means of CFD tools: despite the approximations

linked to numerically solving NS equations this method allows the simulation of three-

dimensional flows on complex geometries, which explains the presence of a vaste literature on

the analysis of radial turbines including comparative studies among different geometrical

configurations. On the other hand “research activities pertaining to the inverse design problem

has not been extensive” (Tjokroaminata, [26]).

Since both three dimensional design and performance analysis (second and third step of the

design process, respectively) are topics of the present work, a review of the main results of

public domain in these fields is now presented.

1.4.1 Inverse problem: the volute

In three articles Whitfield et al. ([12], [31], [32]) give a comprehensive treatment of vaneless

volutes for radial inflow turbines, which is taken as main reference for volute design in this

Page 20: Numerical Modelling of a Radial Inflow Turbine with and

9

thesis. Firstly they developed a practical method, suitable for implementation in a program code,

based on simple hypotheses of incompressible flow and free vortex law. Since the outflow angle

is specified as input parameter the aim of Whitfield’s method is to design distributions of

centroid radius and cross section area which ensure the flow to be delivered uniformly and at the

appropriate angle to the impeller. A subsequent experimental investigation on such obtained

geometry showed that the free vortex pattern is only a fair approximation over the first 180° of

azimuth angle, while as the tongue position is approached the variation of the tangential

component of velocity with radius reduces considerably (Whitfield et al., [32]). The free vortex

law was therefore modified accordingly by means of a so called vortex exponent.

The method gives satisfactory results, at least at first attempt, but relies on several purely

empirical considerations such as the shape of the vortex exponent or the variation of Mach

number with azimuth angle. Moreover no indications are given about the optimal shape of the

volute cross sections. The latter aspect is studied by Shah et al. [20] who, after the comparison

between trapezoidal, Bezier-trapezoidal and circular cross sections, suggested that the circular

cross section will give a better efficiency.

Abidat et al. [1] suggest to design the distribution of centroid radius by means of a Bezier

polynomial and the distribution of cross section area by assuming a linear reduction of mass flow

with azimuth angle. The employment of parametric curves makes this method flexible for the

designer and allows her to easily take into account other possible constraints (for example on the

radius at volute outlet) but has somehow an empirical basis. Moreover there is no careful design

of the critical region around the tongue.

1.4.2 Inverse problem: the rotor

Yang [34] pointed out that the complex nature of the flow through a real turbomachine would

make a three-dimensional design procedure difficult if not impossible: for this reason lot of

effort was put on the development of approximate two-dimensional methods in the hub-to-

shroud plane.

An example of such method is presented in the work by Smith & Hamrick [24]. They prescribed

as input parameters blade shape, hub shape and velocity distribution along the hub, then they

introduced an estimated streamline from inlet to outlet of the rotor and checked for continuity of

flow through the annular streamtube within the hub and the streamline. If continuity is not

established, the streamline spacing is adjusted accordingly and another annular streamtube is

constructed over the first, following the same criterion: in this way the final streamline

determines the shape of the impeller shroud.

This method relies on the assumptions of isentropic, steady and non-viscous flow, but its real

limitation is due to the arbitrary choice of the input parameters, which has no theoretical basis

although strongly affects the final solution.

A turbomachinery blading design method in three-dimensional invscid flow was suggested by

Yang [34]. Here the blade is represented by a sheet of bound vorticity, i.e. bound to the solid

surface of the blade. Under the assumptions of steady inviscid and irrotational flow the only

vorticity in the flow field is that generated on the blade surfaces, which is related to the

circulation on a closed path around the axis of rotation: by carefully specifying the mean swirl

distribution (which, if integrated, gives the total amount of circulation) the distribution of bound

vorticity is specified as well, and the blade surface geometry is obtained as that location of the

bound vortex sheet in which the normal velocity vanishes. The condition of non-penetrating flow

must infact hold on the solid surface of the blade.

Tjokroaminata [26] highlights two main drawbacks of Yang’s method:

Page 21: Numerical Modelling of a Radial Inflow Turbine with and

10

“for a wide variety of swirl distributions there always exists a region of inviscid reversed

flow on the pressure surface of the blade [...] which may result in flow separation”

(Tjokroaminata, [26])

Blades obtained with such design tend to be highly twisted, which may lead to structural

problems especially in applications where the rotational speed of the impeller is high

1.4.3 Inverse problem: the remaining components

Diffuser and fixed stator are relatively simple components and their design is not investigated

extensively.

Siggeirsson & Gunnarsson [22] report that “when the diffusion angle [(defined as the slope of the

wall of the divergent pipe)] is large the diffusion rate is rapid and can cause boundary layer

separation resulting in flow mixing and stagnation pressure losses. On the other hand if the

diffusion rate is too low the required length of the diffuser will be very large and the fluid

friction losses increase”. This opinion is shared by other authors such as Dixon [6], who sets to

7° - 8° the value of the diffusion angle which gives an optimal rate of diffusion.

Khader [9] relied on a direct approach for the design of a fixed-geometry nozzle ring. Straight

symmetric blades with rounded LE and TE are chosen, and an initial 3D geometry of the

component is guessed. A following analysis with CFD is performed, and modifications on the

incidence angle of the blades are made until no separation of the flow is detected. The method

followed by Khader, i.e. the iteration of the analysis step in order to reach an optimized design,

may be rather time consuming, but allows to end up with a configuration in which incidence

losses are minimized.

1.4.4 Direct problem: performance analysis

Simpson [23] performed a CFD analysis of existing test turbine geometries in both vaned and

vaneless configurations. A total of six geometries were analysed and the results compared with

measured turbine performance data. “Steady state predictions showed good agreement with the

experimental trends confirming the vaneless stators to yield higher efficiencies across the full

operating range” [23]

According to the author, vaned stators lead to a higher level of losses because of the wake

detaching from vane trailing edges, boundary layer growth and secondary flows.

Spence et al. [25] tested three pairs of vaneless and corresponding vaned stators within a range of

pressure ratios and flow rates. For each pair of stators the rotor was the same and the operating

conditions were identical. “The vaneless volutes delivered consistent and significant efficiency

advantages over the vaned stators over the complete range of pressure ratios tested. At the

design operating conditions, the efficiency advantage was between 2% and 3.5%” (Spence et al,

[25]).

Padzillah et al. [17] compared nozzled and nozzleless turbines under pulsating flow and found

that “the differences in flow angle distribution between increasing and decreasing pressure

instances during pulsating flow operation is more prominent in the nozzleless volute than in its

nozzled counterpart”, suggesting that the addition of a nozzle ring leads the turbine to more

stable flow angle configurations at off-design points. On the other hand Baines & Lavy [2]

claimed that the advantage of the vaned configuration consists of its highest peak efficiency at

design point, as at off-design this efficiency drops dramatically.

Page 22: Numerical Modelling of a Radial Inflow Turbine with and

11

2 – MOTIVATION AND OBJECTIVE

2.1 Motivation

It is estimated (ATRI, [27]) that fuel-related cost accounts for around 35% of overall operational

costs of trucking, which justifies the need to improve engine efficiency and nowadays’ extensive

research on turbochargers for applications on heavy-duty vehicles.

Despite attempts to find optimal solutions have been done (and have been summarized in the

literature review), the problem of designing the turbine of a turbocharger so to guarantee the best

achievable performance is still open. The reason is that the complexity of the flow within a

turbomachinery requires the introduction of simplifying assumptions about its mean motion,

which may lead the designer to neglect some dynamics in the flow pattern (typically secondary

flows) and misevaluate losses.

Moreover, even assuming an ideal geometrical configuration, this would be valid for just one

design point. During a combustion cycle, however, pressure and temperature of the exhaust

gases vary over a considerable range, thus the performance of that configuration should be

optimal also at off-design points. Variable angle stators try to achieve the latter condition by

continuously adapting the inflow angle at rotor inlet, hence reducing incidence losses, but the

implementation of such devices is by now considered unfeasible due to high mechanical stresses

and vibrations.

A comparative study performed by Mora [14] pointed out that a turbine with fixed nozzle ring

has both higher efficiency and power output with respect to a vaneless turbine throughout all the

combustion cycle. This statement is based on a meanline analysis, but in order to be tested a

three-dimensional design of the two turbine configurations should be developed and an

investigation should be performed by means of CFD techniques so to describe the internal flow

in details and identify sources of losses otherwise undetectable.

2.2 Objective

In light of the previous research carried out by Mora, and considering the existence of different

opinions reported in literature, the purpose of this paper is to present a comparative

performance assessment between two turbine configurations, vaneless and with static nozzle

ring, working under the same operating conditions at both design and off design points.

Page 23: Numerical Modelling of a Radial Inflow Turbine with and

12

3 – METHODOLOGY AND TOOLS

3.1 Methodology

The strategy used in the present work is meant to be an implementation of the design process

which was described in the introductory section. In order to achieve the objective several steps

were followed

Literature review

A literature review was carried out firstly. The purpose was to collect background

information about existing configurations of radial turbines and studies of their

performance, as well as techniques for the design of specific components, when

available.

Analysis of the meanline design

The reference meanline design was analysed in details, with specific focus on identifying

the average flow parameters at inlet and outlet of all components of the turbine and

potential geometrical constraints. This phase defines the inputs which enter the

subsequent step.

3D design of components

A theory-based procedure was developed for the 3D design of the volute and the

impeller. Diffuser and static nozzle ring were not designed with an inverse approach but

as first attempt geometries due to their relatively simple shapes. All the components were

drawn by means of a CAD software (for the rotor a dedicated software, allowing a high

flexibility in the design, was employed).

CFD simulation setup

All components were meshed separately. Structured hexahedral mesh was used for the

rotor blade passage while coarser tetrahedral mesh for the remaining components (with a

refinement close to the walls, so to better describe the boundary layer). After meshing,

vaneless and vaned configurations were assembled: the connection of different meshes

was obtained by specifying suitable interfaces on the contact surfaces. Fluid

thermodynamic properties and boundary conditions were specified as well, and an

appropriate turbulence model was chosen for the solution of RANS. At the end of this

step the two turbine configurations were ready for a CFD steady simulation.

Final design

The results of CFD simulations on both vaneless and vaned cases were compared and the

original 3D design was modified until the average velocity vector of the flow at rotor

inlet was the same for both configurations (since rotor and diffuser are also the same, this

guarantees a “fair” comparison between the two turbines, which only differ by the

presence/absence of a static nozzle ring: every difference pointed out in a subsequent

performance analysis may thus be attributed to that component). At the end of this

iterative phase the final geometrical configurations were obtained and the design process

is concluded.

Analysis of results

Given the two final configurations, comparative performance analysis was done at design

and off-design points in order to identify the “best” one, not only in terms of peak

efficiency but within the whole operative range.

Page 24: Numerical Modelling of a Radial Inflow Turbine with and

13

3.2 Tools – Software

The methodology contains several sub-objectives to be reached before the final comparative

analysis can be done: drawing the 3D model of a component, meshing a blade passage...

For each intermediate task a set of computer programs is employed.

𝑴𝑨𝑻𝑳𝑨𝑩®

MATLAB® is a numerical environment suitable for matrix computation and

implementation of algorithms. Here it is used to implement the 3D design procedure of

the volute and to generate plots from vectors of data.

𝑺𝑶𝑳𝑰𝑫𝑾𝑶𝑹𝑲𝑺®

SOLIDWORKS® is the CAD software which was used to build the complete geometrical

models of volute, nozzle ring and diffuser. It is flexible because it allows to specify

relative constraints between parts of a sketch and adapt it when a modification is done, so

that it is most suitable for the iterative phases of the design process.

Moreover files generated in SOLIDWORKS® can be easily transferred to ANSYS®

package for a subsequent analysis (CFD, FEM...).

𝑫𝑬𝑺𝑰𝑮𝑵𝑴𝑶𝑫𝑬𝑳𝑬𝑹®

DESIGNMODELER® is a CAD software within ANSYS® package. It mainly handles

external geometry models, usually created for manufacturing purpose, and allows

modifications (for example the suppression of non meaningful details) so that the model

is ready for meshing and simulation. DESIGNMODELER® may also be used to draw

geometries from scratch.

In this thesis the program imports geometry files from SOLIDWORKS® and edits them

when needed (by rotating or translating components with respect to the frame of

reference) so that the turbine model is assembled correctly after meshing.

𝑩𝑳𝑨𝑫𝑬𝑮𝑬𝑵®

BLADEGEN® is an ANSYS® software specific for the design of rotative machinery

blades. The designer is allowed to specify the evolution of some representative sections

of the blade in a cylindrical frame of reference and the thickness distribution: the

program automatically generates the CAD model of the machine and monitors several

key parameters such as the cross-section area of the flow channel, the flow angle

distribution...

The design of the impeller of the radial turbine was done with BLADEGEN®.

𝑻𝑼𝑹𝑩𝑶𝑮𝑹𝑰𝑫®

TURBOGRID® is an ANSYS® software, specific for rotating machinery, which creates

high quality hexahedral meshes. The program imports the model of the impeller from

BLADEGEN® and meshes the blade passage: when ATM default feature is enabled, the

program chooses the optimal topology for a given blade geometry and allows to create a

good quality mesh in a highly automated way and with minimal effort, with no need for

control point adjustment.

𝑴𝑬𝑺𝑯𝑰𝑵𝑮®

The tool ANSYS MESHING® imports the geometry of a component from

DESIGNMODELER® and allows the creation of a mesh in a guided and automated way.

The designer has control over width and shape of the cells, can make local refinements

close to sharp edges or narrow passages and can build inflation layers, i.e. structured

Page 25: Numerical Modelling of a Radial Inflow Turbine with and

14

mesh close to a wall to capture the boundary layer. Once fixed the setup parameters,

mesh generation is program-controlled, thus the process is fast and robust, and

subsequent modifications are directly implementable.

Volute, diffuser and nozzle ring are meshed with this software.

𝑪𝑭𝑿®

CFX® is a general-purpose commercial CFD software which is commonly used in

problems involving rotating machinery (turbines, pumps, fans...). CFX® has a

preprocessor in which the setup for the simulation is defined: this includes the definition

of fluid properties, boundary conditions, turbulence model, target of accuracy... All

components of the turbine, once meshed, are imported in CFX® preprocessor and

assembled there to reach the final configuration (either vaneless or vaned).

The final solution is considered achieved when some monitoring parameters, typically

the residuals of mass and momentum equations, have converged below a minimal

threshold value. CFX® also includes a postprocessor which allows to analyse the solution

and obtain a physical interpretation of the result.

Page 26: Numerical Modelling of a Radial Inflow Turbine with and

15

4 – LIMITATIONS

Uncertainties of CFD computation

The present work consists of a numerical performance evaluation of a radial turbine, as

the internal flow is modelled with a CFD commercial software. However, “due to

modelling, discretization, and computation errors, the results obtained from CFD

simulations have a certain level of uncertainty. It is important to understand the sources

of CFD simulation errors and their magnitudes to be able to assess the magnitude of the

uncertainty in the results” [35].

Such uncertainty is mainly linked to the mathematical model describing the flow

(accuracy of the turbulence model, model for Reynolds stress when solving RANS, wall

functions...), to the method for the discretization of PDE’s, to the quality of the

computational grid and computer round-off errors. Moreover the validity of the

numerical solution is jeopardized wherever large regions of separated flow exist in the

domain, which may be the case for certain turbine configurations, especially at off-design

points. These sources of uncertainty constitute a limitation for the study because a certain

set of data (for example a performance difference between vaneless and vaned turbine

configurations at given working condition) may fall inside the uncertainty range, and

since experimental results are not available and cannot be used to validate the numerical

model, those data would lead to misleading interpretations.

Simplifications in the computational domain

In the following chapters it will be seen that the mesh of a full radial turbine may be

rather “heavy”, especially if low values of 𝑦+ are set as requirement. A simplifying

approach present in literature [22] and also adopted in this work is to mesh a single rotor

blade passage (SBP) and assume symmetry of the flow around the rotational axis: in this

way, however, the non-uniformities of the flow within different rotor blade-to-blade

channels are neglected.

Simplifications in the 3D geometrical model

The design of simplified geometrical models may lead to neglecting or misevaluating

certain features of the flow pattern, therefore it constitutes a limitation to the validity of

the results obtained in this study. This is particularly true for the rotor, where tip

clearance, scalloping and other geometrical details associated with the onset of secondary

flows are not accurately modelled at a first-attempt design.

Moreover the design of all components of the turbine is based on modelling assumptions

which introduce further simplifications (details in the following chapter)

Page 27: Numerical Modelling of a Radial Inflow Turbine with and

16

5 – DESIGN OF COMPONENTS

5.1 Design of the volute

In a radial inflow turbine the volute has the purpose to decrease the static pressure of the working

fluid and to increase its speed (conversion to kinetic energy) in order to reach desired values of

velocity and flow angle at rotor inlet. Moreover the volute must distribute the flow uniformly

along the azimuth direction and perform the energy conversion as efficiently as possible, that is

with a minimum loss in stagnation pressure.

With reference to Fig.9, the constraints on the volute design are the following:

Radius at outlet (𝑅2) and passage width (𝑏2). These are set by the rotor geometry and

define the volute discharge area (𝐴2).

Mach number (𝑀2) and absolute flow angle (𝛼2) at outlet. These requirements are

imposed by the designed performance of the rotor.

Thermodynamic flow conditions at volute inlet (𝑃01 , 𝑇01 , γ). These parameters are set by

the working point of the engine.

Mach number (𝑀1) at inlet, linked to the velocity of the exhaust gases.

The goal of the preliminary 3D design is to size the flow passage in terms of the variation, with

azimuth angle (φ), of cross-section area and centroid radius. The solving strategy illustrated in

the following section is a modified version of the procedure whose original development is due

to Whitfield [31], and which is based on the assumptions of adiabatic incompressible flow and

conservation of angular momentum. Notice that incompressibility is a rough assumption, as the

flow undergoes a variation of Mach number inside the volute: however, according to the design

parameters (Tab.2), the flow regime is low subsonic, moreover the resultant design turns out to

be acceptable at first attempt, thus the assumption is ultimately justified by experience.

Figure 9: schematic diagram of a vaneless volute casing (Whitfield, [31])

Page 28: Numerical Modelling of a Radial Inflow Turbine with and

17

5.1.1 Theoretical procedure

Newton’s second law of motion applied to moment forces reads: “For a system of mass m, the

vector sum of the moments of all external forces acting on the system about some arbitrary axis

A–A fixed in space is equal to the time rate of change of angular momentum of the system about

that axis” (Dixon, [6])

𝑀 = 𝑑

𝑑𝑡(𝑅𝐶𝜃) (4.1)

In a purely ideal case no external forces, so moments, are applied on a fluid particle moving from

the volute inlet to the volute outlet. In practice viscous shear forces, even if weak (due to high

temperature of the exhaust gas its viscosity is low), are not absent, and the angular momentum is

not fully conserved. Eqn.(4.1), evaluated from inlet to outlet, then becomes

𝑅2𝐶𝜃2 = 𝑆𝑅1𝐶𝜃1 (4.2)

where S is the angular momentum ratio across the volute.

For an adiabatic flow eqn.(4.2) can be developed in terms of absolute Mach numbers and flow

angles to give the volute radius ratio as

𝑅1

𝑅2=

𝑀2 sin𝛼2

𝑆𝑀1 sin𝛼1 1 + 𝛾 − 1 /2 𝑀1

2

1 + 𝛾 − 1 /2 𝑀22

12

4.3

The area ratio is derived from the conservation of mass between sections 1 and 2 (Fig.9)

𝐴1

𝐴2=𝑃02

𝑃01

𝑀2 cos𝛼2

𝑀1 sin𝛼1 1 + 𝛾 − 1 /2 𝑀2

2

1 + 𝛾 − 1 /2 𝑀12

− 𝛾+1

2 𝛾−1

(4.4)

where the stagnation pressure at outlet 𝑃02 is a function of the target efficiency of the volute (in

ideal case 𝑃02 = 𝑃01).

It should be noticed that eqns.(4.3) and (4.4) depend on 𝛼1, which, in the original procedure

(Whitfield, [31]), is given as input parameter. However the inflow angle at inlet is unknown at

this stage, because it is linked to the orientation of section 1 (take Fig.9 as reference) which in

turn depends on the tangent to the centroid locus in section 1, not yet determined. Thus 𝛼1 must

be either entred as guess parameter or estimated.

Here 𝛼1 is derived by equating the expressions of non-dimensionalised mass flow rate 𝜃 =𝑚

𝜌0𝑎0𝐴

between sections 1 and 2, expressed in terms of Mach numbers

𝑀1 sin𝛼1 1 +𝛾 − 1

2𝑀1

2 − 𝛾+1

2 𝛾−1

= 𝑀2 cos𝛼2 1 +𝛾 − 1

2𝑀2

2 − 𝛾+1

2 𝛾−1

(4.5)

This approach is not rigorous because the quantity which is conserved is not 𝜃 but 𝑚 . However

the final design under such approximation will be shown, after CFD simulation, to fulfill the

requirements at rotor inlet, and the qualitative behavior of 𝑅, 𝐴 and 𝛼 with azimuth angle

forecasted by Whitfield is still respected.

Page 29: Numerical Modelling of a Radial Inflow Turbine with and

18

Once the overall volute geometry is defined in terms of inlet-to-outlet radius and area ratio the

next step is to extend the analysis to the derivation of geometric parameters as function of the

azimuth angle. To this purpose the following hypotheses are introduced:

Mach number at passage centroid increases linearly with θ → 𝑀𝑦 = 𝑀1 + Φ

2𝜋 𝑀2 −

𝑀1

Mass flow rate decreases linearly with θ → 𝑚 𝑦

𝑚 = 1−

Φ

2𝜋 +

𝑚 𝑅

𝑚 , where 𝑚 𝑅 is the flow

rate recirculating and mixing below the volute tongue (𝑚 𝑅 ≈ 5% at first estimate)

Stagnation pressure decreases linearly with θ → 𝑃02 = 𝑃01 − Φ

2𝜋 𝑃01 − 𝑃02

The vortex exponent m, which takes into account the modification of the free vortex

condition, varies with θ → 𝑚 = 𝑚0 − 𝑘𝛷𝑝 , where 𝑚0 = 1 is the exponent at volute

inlet while k and p are experimental constants.

Figure 10: Left – velocity profile across the centre-line of the volute section. Right – variation of

centroid radius at two subsequent azimuth positions (Whitfield, [31], adapted)

The swirling flow is subjected to the free vortex relation (conservation of momentum), corrected

by the vortex exponent which models the presence of small tangential forces arising from the

non ideal volute design

𝐶𝜃𝑅𝑚 = 𝑐𝑜𝑛𝑠𝑡 4.6

Eqn.(4.6) allows to express the variation of tangential velocity between any two angular

locations X and Y separated by a small angle 𝛥𝛷 (see Fig.10 – left)

𝐶𝜃𝑦

𝐶𝜃𝑥= 1−

𝛥𝛷

2𝜋

𝐶1𝑅1 sin 𝛼1

𝐶𝑥𝑅𝑥 sin 𝛼𝑥 1 − 𝑆

𝑆𝑥𝑦

𝑅𝑥𝑅2 𝑚𝑥

𝑅2

𝑅𝑦

𝑚𝑦

(4.7)

where 𝑆𝑥𝑦 represents the local dissipation of angular momentum due to wall skin friction forces.

Page 30: Numerical Modelling of a Radial Inflow Turbine with and

19

The flow angle at Y can now be derived from eqn.(4.7) in terms of the known Mach numbers

and 𝛼𝑥 (known as well): the centroid radius 𝑅𝑦 , unknown at this stage, is derived from

geometrical considerations (Fig.10 – right).

Finally the conservation of mass is applied between section 1 and the generic section Y in order

to find an expression for the area ratio

sin 𝛼𝑦 = 𝑆𝑥𝑦 sin 𝛼𝑥 𝑅𝑥𝑅2 𝑚𝑥

𝑅2

𝑅𝑦

𝑚𝑦

1 + 𝛾 − 1 /2 𝑀𝑦

2

1 + 𝛾 − 1 /2 𝑀𝑥2

12

(4.8)

𝑅𝑦 = 𝑅𝑥 − 𝛥𝑅𝑥𝑦 = 𝑅𝑥 −𝑅𝑦𝛥𝛷

tan𝛼𝑥 → 𝑅𝑦 =

𝑅𝑥

1 +𝛥𝛷

tan 𝛼𝑥

(4.9)

𝐴1

𝐴𝑦=𝑃0𝑦

𝑃01

𝑚

𝑚 𝑦

𝑀𝑦 sin𝛼𝑦

𝑀1 sin𝛼1 1 + 𝛾 − 1 /2 𝑀𝑦

2

1 + 𝛾 − 1 /2 𝑀12

− 𝛾+1

2 𝛾−1

(4.10)

5.1.2 Implementation of the theoretical procedure

The procedure described above was implemented in a MATLAB®code, with the input parameters

coming from the meanline design.

In order to compare vaneless and vaned volutes at equal working conditions, both configurations

are required, at design point, to deliver to the rotor a flow with the same average velocity vector.

Thus the design specifications at volute inlet and outlet must be the same as well. However in the

vaned case the nozzle ring behaves as a convergent duct, hence the total acceleration is

distributed between the two components.

PARAMETER VALUE

Radius at nozzle inlet 𝑟𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 = 52.3 [mm]

Radius at nozzle outlet 𝑟𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡 = 41.8 → 41 [mm] (*)

Blade height 𝑕 = 13 [mm]

Table 1: meanline design parameters for nozzle ring

(*) The radius at outlet was slightly changed from the meanline value so to have complete accordance between the

geometrical dimensions of the two volute configurations.

The acceleration that the flow undergoes inside the nozzle ring is derived from the conservation

of mass between inlet and outlet sections. This estimate is approximated, as density and

temperature of the flow are supposed not to change during the process.

𝑚 𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 = 𝑚 𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡 → 𝜌𝑀𝑎 𝑢

2𝜋𝑟𝑕 𝐴

𝑛𝑟−𝑖𝑛𝑙𝑒𝑡

= 𝜌𝑀𝑎 𝑢

2𝜋𝑟𝑕 𝐴

𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡

(4.11)

If the blade height is constant, a rearrangement of eqn.(4.11) leads to an expression for the Mach

number at nozzle inlet, which is the only unknown.

Page 31: Numerical Modelling of a Radial Inflow Turbine with and

20

𝑀𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 = 𝑀𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡

𝑟𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡𝑟𝑛𝑟−𝑖𝑛𝑙𝑒𝑡

(4.12)

where 𝑟𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 and 𝑟𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡 are reported in Tab.1. The input parameters for volute design are

summarized in Tab.2. The solution is shown in Fig.11.

PARAMETER VANELESS VANED (no nozzle)

Mach at volute inlet 𝑀1 = 0.21 𝑀1 = 0.21 Mach at volute outlet 𝑀2 = 0.56 𝑀2 = 𝑀𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 = 0.44

Flow angle at volute outlet 𝛼2 = 68 [°] 𝛼2 = 68 [°] Radius at volute outlet 𝑅2 = 41 [𝑚𝑚] 𝑅2 = 𝑟𝑛𝑟−𝑖𝑛𝑙𝑒𝑡

= 52.3 [𝑚𝑚] Outlet passage width 𝑕 = 13 [𝑚𝑚] 𝑕 = 13 [𝑚𝑚]

Flow properties at inlet

𝑃01 = 2.56 [𝑏𝑎𝑟] 𝑇01 = 846 [𝐾]

𝑚 1 = 0.3322 [𝑘𝑔

𝑠]

𝛾 = 𝛾𝑎𝑖𝑟 |𝑇01= 1.34

𝜌1 = 1.037 [𝑘𝑔

𝑚3]

𝑃01 = 2.56 [𝑏𝑎𝑟] 𝑇01 = 846 [𝐾]

𝑚 1 = 0.3322 [𝑘𝑔

𝑠]

𝛾 = 𝛾𝑎𝑖𝑟 |𝑇01= 1.34

𝜌1 = 1.037 [𝑘𝑔

𝑚3]

Additional constraints 𝑅1 = 82.14 [𝑚𝑚] 𝐴1 = 2733 [𝑚𝑚^2]

𝑅1 = 82.14 [𝑚𝑚] 𝐴1 = 2733 [𝑚𝑚^2]

Table 2: meanline design parameters for vaneless and vaned volutes

Figure 11: theoretical distribution of centroid radius, cross section area and flow angle with

azimuth location

Page 32: Numerical Modelling of a Radial Inflow Turbine with and

21

5.1.3 Implementation under additional constraints

The solution deriving from the application of the theoretical procedure may not be

implementable due to the presence of additional constraints on the design. In particular,

requirements on the maximum size of the turbine may limit radius or area at volute inlet. The

designer may then look for a sub-optimal solution, i.e. a solution which is as close as possible to

the ideal one yet fulfilling all the constraints.

For the present case the meanline design reports values of 𝑅1 and 𝐴1 which were interpreted as

additional geometrical constraints (Tab.2): however the theoretical solution does not match the

values expected at volute inlet, as seen in Fig.11.

Qualitatively, a constraint on 𝑅1 is reflected into a modification of the radius distribution with

respect to the ideal case. Following a strategy suggested by Abidat [1], the modified centroid

radius distribution is modeled with a third order Bezier polynomial, a parametric curve defined

by 4 monitoring points, conveniently chosen, which allow high control over the shape of the

curve and its derivative. Details about the Bezier approach will be given in Section 5.3 “Design

of the Rotor”. The polynomial is numerically derived in order to obtain the flow angle

distribution, which is geometrically represented by the local tangent to the centroid locus (see

Fig.10- right). Given 𝛼 at the generic azimuth position and provided that all the other parameters

remain constant, the distribution of cross section area is computed from eqn.(4.10).

The implementation of the solving procedure under constraints is done in a MATLAB® code and

results are reported in Fig.12 and Fig.13.

Figure 12: distribution of centroid radius, cross section area and flow angle with azimuth

location. Comparison between theoretical and implemented solution, vaneless case

Page 33: Numerical Modelling of a Radial Inflow Turbine with and

22

Figure 13: distribution of centroid radius, cross section area and flow angle with azimuth

location. Comparison between theoretical and implemented solution, vaned case

Observations

Overall the implemented solution follows the theoretical one in the central range of

azimuth positions (from 𝜑 ≈ 100° to 𝜑 ≈ 300° ). Major deviations occur close to the

inlet, where the constraints are specified, but local modifications are done after a

subsequent CFD analysis as part of the iterative phase of the design process.

Fig.13 shows that a modified radius distribution may lead to a non negligible variation of

the outflow angle with respect to the designed value. One solution is to use a higher

degree Bezier polynomial, with more control points in order to model the slope of the

curve at the edges. However this example highlights the difficulty to fulfill all constraints

at once and the limits of inverse design methods.

The Bezier polynomial is transformed into a cartesian curve in space and imported in

SOLIDWORKS®. The curve represents the locus of the centroids of all the cross sections.

Circular cross section is chosen, as it is the simplest shape and leads to an efficient configuration

according to literature (Shah, [20]). The area of each section is sized according to the solution of

the 3D design procedure: slight modifications are made so that, for a correct matching of the

components, the outlet passage width is kept constant at all azimuth positions and equal to the

designed blade height of the impeller,

3D CAD models of the volute casing are shown below (Fig.14).

Page 34: Numerical Modelling of a Radial Inflow Turbine with and

23

Figure 14: 3D geometrical model of the volute casing. Left: vaneless. Right: vaned (no nozzle)

5.2 Design of the nozzle ring

The vaned stator must guarantee the designed flow angle at rotor inlet, ideally without

introducing additional losses.

The constraints on the design of the nozzle ring are below (their values are reported in Tab.1)

Radius at inlet (𝑟𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 ) and radius at outlet (𝑟𝑛𝑟−𝑜𝑢𝑡𝑙𝑒𝑡 ). These must match the outlet

radius of the volute and the inlet radius of the impeller respectively

Blade height (𝑕). This requirement is set by the volute outlet passage width

Moreover, other parameters from the meanline design are specified: mass flow rate from the

volute outlet section (𝑚 = 0.3322 [𝑘𝑔

𝑠]), fluid density (𝜌 = 1.037 [

𝑘𝑔

𝑚3]), speed of sound (𝑎 =

564.3 [𝑚

𝑠]), required flow angle at rotor inlet (𝛼 = 68 [°]).

Given the set of design parameters, a final configuration which fulfills all of them may not be

implementable if chocking occurs inside the nozzle vanes. Thus the condition 𝑀 ≤ 1 must be

verified for the design to be meaningful.

From Fig.15 𝛼 is expressed in terms of nozzle pitch and nozzle throat length as

𝛼 = cos−1 𝑜

𝑠 = cos−1

𝑚

𝜌𝑀𝑎 𝑕𝑍𝑣2𝜋𝑟𝑛𝑟 −𝑖𝑛𝑙𝑒𝑡

𝑍𝑣

(4.13)

where 𝑠 =2𝜋𝑟𝑛𝑟 −𝑖𝑛𝑙𝑒𝑡

𝑍𝑣 is derived from geometrical considerations while 𝑜 =

𝑚

𝜌𝑀𝑎 𝑕𝑍𝑣 expresses the

link between throat area and mass flow rate for a nozzle vane.

Page 35: Numerical Modelling of a Radial Inflow Turbine with and

24

Figure 15: nozzle vane geometry definition (Rajoo & Martinez-Botas, [18])

Eqn.(4.13) allows to compute the Mach number. By substituting the design parameters it is

found 𝑀 ≈ 0.36, thus chocking is avoided at design point (notice there is no guarantee that

𝑀 < 1 is also verified at off-design conditions)

Moreover eqn.(4.13) shows that the number of vanes has no influence on the outflow angle:

according to Rajoo & Martinez-Botas 𝑍𝑣 is the value which optimizes the pitch/chord ratio so to

achieve “compromise between friction losses and good flow guidance” [18]. Meanline design

specifies 𝑍𝑣 = 14.

Figure 16: nozzle ring. Left: sketch in the frontal plane. Right: shape of the blade

For the 3D design of the nozzle ring iterative procedures based on CFD analysis have been found

in literature [9] and the same approach is followed here. The blade has a simple symmetric

profile with straight sides closed at LE and TE by circumference arcs (Fig.16, right). The

Page 36: Numerical Modelling of a Radial Inflow Turbine with and

25

inclination angle (Fig.16, left) is set arbitrarily: as initial guess value the designer chooses the

outlet flow angle 𝛼, since the flow is expected to “follow the blade” in ideal design. This value is

modified iteratively until no separation is detected.

3D CAD model of the nozzle ring is generated in SOLIDWORKS® and is shown in Fig.17.

Figure 17: 3D geometrical model of the nozzle ring

5.3 Design of the rotor

The rotor is the main component of a radial turbine, being the only one which extracts work from

the flow. Apart from the requirement of matching with the volute, which sets the dimensions of

the diameter and the height of the blade at inlet, there are no constraints on the design of the

rotor: the goal is to minimize all sources of losses so to achieve the maximum power output

under designed flow conditions.

5.3.1 Preliminary design

As illustrated in the introductory chapter, nominal design condition is associated with a

minimum in incidence losses. Therefore the rotor blade must be designed so that for all spanwise

locations (from hub to shroud, see Fig.18) the flow enters radially (𝛽2 = 0) and leaves the rotor

axially (𝛼3 = 0).

In literature [21] it is reported that the condition 𝛽2 = 0 actually leads to flow recirculation at the

suction surface of the blade, and the optimal inlet flow angle is identified within the range

𝛽2 = −10° and 𝛽2 = −40°. Given the flow parameters at volute outlet (𝑀2 = 0.56, 𝛼2 = 68°)

and the designed rotational speed (𝑛 = 85000 rev/min) it is 𝛽2 ≈ −26°, which falls inside the

optimal interval. It should be noticed that in general 𝛽2 may differ from the physical angle of

incidence, and this occurs when the blade angle at rotor inlet, namely 𝛽2𝑏 , is non zero. However

strength limitations require the blade to be radial in the segment furthest from the rotational axis,

Page 37: Numerical Modelling of a Radial Inflow Turbine with and

26

where centrifugal forces are dominant: this constraints the inlet blade angle to be zero. In this

configuration the angle of incidence coincides with 𝛽2 thus it is in the optimal range as well.

Figure 18: sketch of the velocity triangles at rotor inlet and outlet (Saravanamuttoo, [19],

adapted)

At rotor outlet the condition 𝛼3 = 0 implies different values of 𝛽3 spanwise along the blade,

because while 𝑤3 is supposed to be constant and equal to its meanline value 𝑤3 = 288 𝑚/𝑠 (this

assumption is only a rough approximation) 𝑈3 increases proportionally with the distance from

the axis of rotation.

SPANWISE POSITION [mm] BLADE ANGLE AT OUTLET [°] (*)

𝑟3 = 11.1 (hub) 𝛽3 = 20.1

𝑟3 = 16.3 𝛽3 = 30.2 𝑟3 = 21.6 𝛽3 = 41.9 𝑟3 = 26.8 𝛽3 = 55.9

𝑟3 = 32 (shroud) 𝛽3 = 81.4

Table 3: relative angle at rotor outlet under nominal design condition

(*) These values are not to be interpreted as hard constraints for the final design but as an approximate guideline,

since they are obtained from simplifying hypotheses based on meanline approach. This is particularly true for hub

and shroud, i.e. close to the walls, where 𝑤3 differs considerably from its meanline value.

Tab.3 reports values of 𝛽3 computed for 5 representative spanwise positions. In ideal case the

blade is designed to be a streamline: indeed if the flow follows the blade “smoothly” incidence

losses are low. This implies that also at rotor outlet 𝛽3 ≡ 𝛽3𝑏 .

So far the blade angles at inlet and outlet have been derived from preliminary design

considerations, but this is not sufficient for the 3D design of the rotor as no indications are given

about the blade angle distribution in the streamwise direction, i.e. along 𝑚 (see Fig.19).

Moreover the evolution of the blade in the θ-direction (namely the “wrap angle distribution”) is

Page 38: Numerical Modelling of a Radial Inflow Turbine with and

27

also undetermined. In the present work the approach to the 3D design consists of modelling the

wrap angle distribution at several spanwise locations by means of parametric Bezier curves.

Figure 19: rotor views. Left: r-z (or meridional) plane. Right: θ-z (or blade to blade) plane

5.3.2 The Bezier curve

Given 𝑛 + 1 control points 𝑃0, 𝑃1,..., 𝑃𝑛 the Bezier curve of degree 𝑛 in space is defined as the

function F: [0,1] → ℝ3 such that

𝐹 𝑡 = 𝑛

𝑖

𝑛

𝑖=0

𝑃𝑖 1− 𝑡 𝑛−𝑖𝑡𝑖 , 𝑡 ∈ 0,1 (4.14)

Bezier curves are interpolation functions. Eqn.(4.14) can be re-arranged so to highlight the set of

functions 𝐵0,𝑛 𝑡 ,𝐵1,𝑛 𝑡 ,… ,𝐵𝑖 ,𝑛 𝑡 ,… ,𝐵𝑛 ,𝑛 𝑡 which represents a basis of the vector space of

all polynomials of degree ≤ 𝑛, also named Bernstein basis.

𝐹 𝑡 = 𝑃𝑖 𝐵𝑖,𝑛(𝑡)

𝑛

𝑖=0

, 𝐵𝑖,𝑛 𝑡 = 𝑛

𝑖 𝑃𝑖 1− 𝑡

𝑛−𝑖𝑡𝑖 (4.15)

Control points 𝑃𝑖 are also the coefficients multiplying each element of the basis; thus by varying

the set of control points the whole vector space can be spanned. Moreover the position of the

control points identify a region inside which the curve will develop. 𝑃0 and 𝑃𝑛 are respectively

the first and the last element of the curve, while the remaining points do no lay on the curve

(apart from the trivial case in which all points are aligned) but locally affect its slope and

curvature.

The advantage to use Bezier curves as design tools is not only due to their flexibility but also to

the fact that the interpolation by means of Bernstein’s polynomials, unlike linear interpolation,

generates “smooth” curves i.e. curves which are continuous also in the first derivative. This

property is important when modelling rotor blades, in which sharp edges must be avoided so to

have continuous variation of the flow angle and limit incidence losses and the risk of local flow

separation.

Page 39: Numerical Modelling of a Radial Inflow Turbine with and

28

As an example Fig.20 shows a Bezier curve of 3rd

order. Bernstein basis is formed by 3rd

order

polynomials (Fig.20, left). As mentioned, the curve is confined within the region identified by

the control points, whose position affects the local slope (Fig.20, right).

Figure 20: Bezier curve of degree 3. Left: basis of vector space. Right: control points (Floater,

[7], adapted)

5.3.3 Implementation of the design strategy

Data relative to the general dimensions of the rotor are provided by the meanline design. The

value of these parameters is reported in Tab.4.

PARAMETER VALUE

Rotational speed (rpm) 𝑛 = 85000

Inlet tip radius 𝑟2 = 40 [𝑚𝑚] Inlet blade height 𝑏2 = 13 [𝑚𝑚]

Exit tip radius 𝑟3𝑡 = 32 [𝑚𝑚]

Exit hub radius 𝑟3𝑕 = 11.1 [𝑚𝑚] Axial length 𝑘𝑎 = 23.9 [𝑚𝑚]

Number of blades 𝑍𝑏 = 12

Table 4: meanline design parameters for the rotor

Five equidistant spanwise locations, namely layers, are identified (the 1st being the hub, the 5

th

being the shroud) and for each of them the wrap angle distribution is modelled by means of a

Bezier curve. This operation is done efficiently in BLADEGEN®, where it is possible to drag, add

or delete control points so to determine the Bezier curve in terms of order and shape.

For the specific case a number of control points between 26 and 31 was used for each layer. High

degree polynomials were chosen in order to locally have control over the curvature, which is not

constant streamwise (it was mentioned that for mechanical reasons the blade is required to

develop radially in its first segment).

Given a wrap angle distribution BLADEGEN® numerically evaluates the tangent at each point.

The resultant plot corresponds to the flow angle distribution, since 𝜃 and 𝛽 are connected by the

relation 𝛽 =𝑑𝜃

𝑑𝑚 (see Fig.19 – right): hence the 𝜃-distribution is designed so that the associated 𝛽

fulfils the condition of axial flow at rotor outlet (Tab.3).

Page 40: Numerical Modelling of a Radial Inflow Turbine with and

29

Figure 21: Rotor: θ-distribution (top), β-distribution (middle), thickness distribution (down)

Page 41: Numerical Modelling of a Radial Inflow Turbine with and

30

For each layer BLADEGEN® must create the blade profile. The latter is defined by a mean line,

which is the wrap angle line, and a streamwise distribution of thickness, which is modelled by

the designer with a Bezier curve. Profiles of adjacent layers are connected together by

streamwise lofting and the result is the generation of the blade surface.

Distributions of 𝜃, 𝛽 and thickness are shown in Fig.21, while the result of such design on the

meridional plane is reported in Fig.22.

Figure 22: Rotor: distribution of wrap angle (top left), flow angle (top right) and thickness

(bottom) in the meridional plane

As seen in the introductory chapter, in order to increase the work exchange the flow must

accelerate inside the rotor, i.e. 𝑤3 > 𝑤2, which implies that for subsonic flows the channel

between rotor blades is convergent. Fig.23 – left shows that this is the case for the present

design, as the cross section area of the channel reduces from inlet to outlet.

The 3D CAD model of the impeller, which results from the design above, is presented in Fig.24.

Page 42: Numerical Modelling of a Radial Inflow Turbine with and

31

Figure 23: Rotor: variation from inlet to outlet of channel cross-section area (left) and lean

angle (right)

Figure 24: 3D geometrical model of the rotor

5.3.4 Supplementary issues on 3D design of the rotor

So far the design of the impeller was made from a geometrical point of view, without due

consideration of the influence on the internal flow pattern. This relationship is impossible to

investigate analytically because it will be seen that, no matter how “good” the design is, the

presence of secondary flows cannot be avoided, thus the problem is fully 3D. CFD analysis is the

most common tool to have an insight on the internal flow, because the only way to have a

comprehensive description of it is by numerically solving the full set of NS equations: however

in this section an analytical approach is presented, whose purpose is to highlight the influence of

some design choices on the development of secondary flows, and possibly suggest improvements

at subsequent phases of the deign process.

Page 43: Numerical Modelling of a Radial Inflow Turbine with and

32

This analysis is an adaptation of the study presented by Van den Braembussche [28] and

originally performed on the impeller of a radial compressor.

Blade lean is defined as the variation of 𝜃 of the blade from hub to shroud. Under the assumption

that the blade lean is null, i.e. 𝜕𝜃

𝜕𝑛= 0 (this is almost satisfied, as the lean angle lies within the

interval [0°,4°] everywhere from inlet to outlet, see Fig.23 – right) the components of velocity

which induce centrifugal accelerations in the meridional plane are:

Meridional velocity 𝑊𝑚 , whose radius of curvature is the one of the streamline, namely 𝑅𝑛 (the

subscript 𝑛 denotes that this radius is aligned with the spanwise direction)

Tangential velocity 𝑉𝑡 = 𝑊𝑡 − 𝛺𝑅, whose radius of curvature is 𝑅. Notice that this vector is

orthogonal to the meridional plane, but the resultant acceleration is in plane

Figure 25: geometrical definition of the problem

At all streamwise sections the overall centrifugal acceleration is balanced by a pressure gradient

orthogonal to the streamsurface (Fig.25). The equilibrium is expressed by eqn.(4.16)

1

𝜌

𝜕𝑝

𝜕𝑛= 𝑊𝑢 − 𝛺𝑅

2

𝑅cos 𝜆 −

𝑊𝑚2

𝑅𝑛 (4.16)

where 𝜆 is the angle between the meridional component of the streamline and the axis of

rotation. At inlet cos λ = 0 and the acceleration due to the curvature of the streamline increases

from hub to shroud because of decreasing 𝑅𝑛 , hence the pressure gradient decreases. At outlet

the direction of the pressure gradient depends on the values of 𝑉𝑡 and 𝑊𝑚, but with increasing 𝑅

from hub to shroud the spanwise acceleration due to the tangential velocity component tends to increase

as well (𝑎𝑡 ∝ 𝛺2𝑅), thus the pressure gradient is still negative.

The above considerations suggest that, together with the main motion of the flow from inlet to

outlet, there exist also a secondary flow moving from hub to shroud, i.e. against the spanwise

pressure gradient.

Imposing 𝜕𝑝

𝜕𝑛= 0 eqn.(4.16) gives the value of 𝑅𝑛 for zero pressure gradient

Page 44: Numerical Modelling of a Radial Inflow Turbine with and

33

𝑅𝑛 =𝑊𝑚

2𝑅

𝑉𝑡2 cos 𝜆

(4.17)

However 𝑊𝑚 is higher at the suction side of the blade (where the flow is accelerated) and lower

at the pressure side, and from eqn.(4.17) two different values of 𝑅𝑛 should exist at the same

point, thus there is no design which can eliminate the spanwise pressure gradient, and the

correspondent secondary flow, at both sides of the blade. The hub-to-shroud pressure gradient

can be reduced by increasing the curvature radius of the meridional contour or by reducing the

blade height (hub-to-shroud distance).

Moreover there also exists a pressure gradient in the blade-to-blade plane because of the flow

moving between pressure side and suction side of adjacent blades. Inside the channel the flow

feels the simultaneous effect of both pressure gradients. The situation is depicted in Fig.26.

Figure 26: pressure distribution in a crosswise section. Effect of spanwise pressure gradient

(left), effect of blade-to-blade pressure gradient (middle), ensemble (right) (Van den

Braembussche, [28])

A possible strategy to reduce the effect of the secondary flow is the employment of splitter

blades.

5.4 Design of the diffuser

The diffuser must convert part of the kinetic energy of the flow into pressure in order to reach

the condition 𝑃4 > 𝑃𝑎𝑡𝑚 at the outlet of the turbine and allow the flow to be discharged.

The constraints on the design of the diffuser are

Inner radius (𝑟3𝑕) at diffuser inlet, which must be equal to the hub radius at rotor outlet

Outer radius (𝑟3𝑡) at diffuser inlet, which must be equal to the shroud radius at rotor outlet

Values provided by the meanline design are reported in Tab.5

PARAMETER VALUE

Outer radius at diffuser outlet 𝑟4𝑡 = 38.6 [mm]

Inner radius at diffuser outlet 𝑟4𝑕 = 0 [mm]

Axial length 𝑙 = 38.6 [mm]

Table 5: meanline design parameters for the diffuser

The main problem when designing the diffusers is the risk of boundary layer separation at the

wall [8], which depends on the diffusion angle ψ (Fig.27 – left).

Page 45: Numerical Modelling of a Radial Inflow Turbine with and

34

Figure 27: conical diffuser. Left: 2D sketch. Right: lines of appreciable stall for given

geometrical configuration (Blevins, [4], adapted)

From meanline parameters the non-dimensional length of the diffuser is 𝑙

𝑟3𝑡= 1.21 while the

designed diffusion angle is expressed from Fig.27 – left as tan𝜓 = 𝑟4𝑡−𝑟3𝑡

𝑙 𝑦𝑖𝑒𝑙𝑑𝑠 2𝜓 ≈ 19.4°.

An extrapolation from the graph in Fig.27 – right suggests that this point is below the limit of

appreciable stall for conical diffusers, hence the present design configuration can be

implemented.

The 3D CAD model of the diffuser is presented below (Fig.28)

Figure 28: 3D geometrical model of the diffuser

Page 46: Numerical Modelling of a Radial Inflow Turbine with and

35

6 – MESH GENERATION

The problem of how to partition the flow domain in order to create a grid on which discretized

NS equations can be solved is of primary importance in CFD, because it directly affects the

quality of the simulation and the computational time. From a practical point of view the ideal

mesh should be able to capture the critical aspects of the flow (presence of regions of separated

flow, recirculation bubbles, evaluation of losses...) with simulations lasting at most a few hours,

which implies that the mesh must be refined only in regions in which it is needed (typically in

the boundary layer, or where high velocity gradients are expected).

The first part of this section focuses on theoretical aspects about mesh generation: how to choose

an appropriate mesh for the problem in exam, how to mesh regions of the domain close to a wall

and the techniques to check if the mesh is “good”. In the second part the solution which was

implemented is illustrated and the mesh quality is assessed.

6.1 Choice of the grid

Two different techniques are employed for the discretization of a geometric domain. In the so

called structured grid the cells are ordered in a I×J×K array so that given whatever grid point

inside the domain this is univocally identified by a set of coordinates, say 𝑃𝑖 ,𝑗 ,𝑘 , and the points in

its neighbourhood are implicitly known (they will be 𝑃𝑖 ,𝑗 ,𝑘−1, 𝑃𝑖 ,𝑗 ,𝑘+1...). This implies that there

exist a regular pattern of connections among grid points which are close to each other. On the

other hand in an unsctuctured grid the above regularity is not present hence neighbouring cells

cannot be directly accessed by their indexes. The different way to build the two grids also affects

the geometrical shape of their cells, i.e. the elements: hexahedra are usually employed in

structured meshes while unstructured meshes are formed by tetrahedral elements or

combinations of different solids (see Fig.29).

Figure 29: elements of a 3D mesh - tetrahedron, hexahedron, prism, pyramid

The choice of the type of mesh should be done according to the following considerations:

Unstructured meshes can easily model every kind of domain because the shape of the

elements which is employed is not constrained to hexahedra. For the same reason also the

element size can vary considerably between adjacent cells. This flexibility is needed

when the geometry to be meshed is complex or when fast variation in the grid spacing is

desirable (for example close to the walls).

For the same amount of cells structured grids based on hexahedra allow the highest

accuracy in the solution. On the contrary unstructured grids tend to generate more

skewed elements, with consequent numerical errors.

The generation of an unstructured grid is much faster than a structured one. The time

strongly depends on the complexity of the problem, but while for the former it is usually

in the order of hours for the latter it can take up to weeks (Khare et al., [10]).

Page 47: Numerical Modelling of a Radial Inflow Turbine with and

36

Meshing a radial turbine is a challenging task because structured hexahedral mesh would be

desirable for high computational accuracy, especially in regions where losses most probably

occur, but the complexity of the 3D geometry makes the unstructured mesh approach more

suitable. In the present work a hybrid solution is suggested. The mesh is structured in the

impeller because it was seen that complex secondary flows occur inside the blade-to-blade

channels and their accurate evaluation is critical for a correct performance assessment of the

turbine. Achieving such a degree of precision is less crucial for other components, like the volute

or the diffuser, where instead the main source of losses is given by skin friction: for this reason

an unstructured mesh is used, which becomes more regular only close to the walls in order to

model accurately the boundary layer.

6.2 Meshing the boundary layer

Boundary layer is a thin region adjacent to a solid surface in which there exist a strong velocity

gradient orthogonal to the surface itself. The reason is that mechanical equilibrium is achieved

between molecules of the flow and the wall in the contact region so that as a macroscopic effect

the velocities of the flow and the solid surface are equal (this constraint is called no slip

condition). The existence of a sharp velocity gradient suggests that for a “good description” of

the flow the grid inside the boundary layer shall be refined. But what happens qualitatively with

increasing Reynolds number is that the region where viscous effects are relevant gets more and

more confined to the walls and the boundary layer becomes thinner, thus the grid resolution must

increase accordingly, leading to an increase of the computational effort.

A rough evaluation of Reynolds number can be done taking as reference values the ones at

turbine inlet: 𝑈𝑟𝑒𝑓 is the meanline velocity, 𝐿𝑟𝑒𝑓 is the radius of the duct at volute inlet and

𝜈|𝑇=𝑇𝑟𝑒𝑓 is the kinematic viscosity of air at 𝑇 = 𝑇𝑖𝑛𝑙𝑒𝑡 at design point.

𝑅𝑒 =𝑈𝑟𝑒𝑓 𝐿𝑟𝑒𝑓

𝜈|𝑇=𝑇𝑟𝑒𝑓

≈120

𝑚

𝑠 ∗ 30 ∗ 10−3[𝑚]

9.06 ∗ 10−5 𝑚2

𝑠

≈ 4 ∗ 104 (5.1)

Eqn.(5.1) shows that 𝑅𝑒 is high, in the order of 104. The flow regime is turbulent, and in the

boundary layer exchange of momentum takes place not only between adjacent layers, at

molecular scale, but together with an exchange of fluid particles.

Figure 30: velocity profile in a turbulent boundary layer (Bakker, [3])

Page 48: Numerical Modelling of a Radial Inflow Turbine with and

37

As a consequence the region close to the wall is subject to steep gradients normal to the

boundary (Fig.30 illustrates a typical velocity profile) and directly resolving the flow with a

suitable mesh is computationally demanding.

Experimental investigation showed that the boundary layer can be divided into an inner and an

outer region. The first region is dominated by viscous effects, and the velocity is a function of

the coordinate 𝑦+ which represents the distance from the wall nondimensionalized by the viscous

scale 𝑢∗

𝜈 (𝑢∗ ≝

𝜏0

𝜌 is called friction velocity).

𝑈

𝑢∗= 𝑓

𝑦𝑢∗𝜈 = 𝑓 𝑦+ (5.2)

The second region is dominated by turbulent mixing and the flow seems not to feel the presence

of the wall. The difference of velocity with respect to the reference value (called velocity defect)

is function of the coordinate 𝜉 which represents the distance from the wall nondimensionalized

by the boundary layer thickness 𝛿.

𝑈 − 𝑈𝑟𝑒𝑓

𝑢∗= 𝑓

𝑦

𝛿 = 𝑓 𝜉 (5.3)

The distance from the wall, y, scales differently in the two regions. However there exist an

intermediate overlap region in which the two expressions (eqn.(5.2) & eqn.(5.3)) are both valid.

By equating their derivatives (for details see Kundu, [11]) it is proved that the overlap region is

described by a logarithmic law, and the so called logarithmic layer becomes wider with

increasing 𝑅𝑒. The situation is illustrated in Fig.31.

Figure 31: non-dimensional velocity as function of 𝑦+ in the inner region (Kundu, [11])

Page 49: Numerical Modelling of a Radial Inflow Turbine with and

38

Under the assumption that the logarithmic behaviour can be used to model the velocity

distribution near the wall, this provides a suitable law to link distance from the wall to velocity,

hence allows the estimation of the flow shear stress. In this way it is not necessary to resolve the

boundary layer because the velocity in the inner region can be estimated through the log-law,

thus a coarser mesh can be used.

Summarizing, there are two approaches to model the flow in the near-wall region

Wall function method. It is based on empirical formulas which link the velocity of the

flow to the position in the inner region of the boundary layer, thus avoiding to resolve it

and saving computational time. However additional assumptions must be introduced in

order to justify the validity of the wall function in the whole inner region, and this may

lower accuracy of the results.

Low-Reynolds-Number method. This method resolves the details of the boundary layer

profile by using very fine meshes with the first node located at 𝑦+ ∼ 1 or even closer to

the wall.

6.3 Quality of the mesh

There exist many criteria for evaluating the quality of a mesh, and an in-depth discussion on this

topic is beyond the purpose of the present work. However from a practical point of view a mesh

is considered of good quality if its elements are not warped too much with respect to their

nominal shape (tetrahedron, pyramid...): for example if an element is too stretched in a certain

direction (see Fig.32) the variation of any flow characteristic in that direction will be detected

with less accuracy because the grid behaves as if it were coarser.

Figure 32: stretching of a quadrilateral element. Nominal shape (left), deformed shape (right)

In this section the following quality parameters are discussed:

6.3.1 Skewness

Skewness is the measure of how close the shape of a cell is from the ideal shape. A possible way

to calculate it is through the so called normalized angle deviation method

𝑠𝑘𝑒𝑤𝑛𝑒𝑠𝑠 ≝ 𝑚𝑎𝑥 𝜃𝑚𝑎𝑥 − 𝜃𝑒

180 − 𝜃𝑚𝑎𝑥,𝜃𝑒 − 𝜃𝑚𝑖𝑛

𝜃𝑒 (5.4)

Eqn.(5.4) evaluates the deviations of the maximum (𝜃𝑚𝑎𝑥 ) and the minimum (𝜃𝑚𝑖𝑛 ) angle with

respect to the angle relative to an equiangular cell (𝜃𝑒 , which represents the ideal case) and takes

as value for the skewness the maximum between both.

A value of 0 represents an equilateral cell while a value of 1 stands for a degenerated cell (for

example in the case of a 2D cell this would become a 1D segment). Skewness is considered good

for values up to 0.5.

Page 50: Numerical Modelling of a Radial Inflow Turbine with and

39

6.3.2 Orthogonal quality

Orthogonal quality is another way to evaluate how a cell is close to its ideal shape. Taking the

2D cell in Fig.33 as reference, orthogonal quality is the minimum of eqn.(5.5) computed ∀i

𝑂𝑄 ≝𝐴𝑖 . 𝑒𝑖

𝐴𝑖 𝑒𝑖 (5.5)

where 𝑒𝑖 is the vector joining the centroid of the cell with the centroid of the edge and 𝐴𝑖 is the

vector normal to the edge. Eqn.(5.5) is a measure of how much 𝑒𝑖 and 𝐴𝑖 are aligned, because the

scalar product at the numerator depends on the cosine of the angle between the two vectors.

The range for the orthogonal quality is [0,1], and the closer to 1 the more equilateral is the cell

(in this case, infact, all the vectors are aligned).

Figure 33: orthogonal quality on a 2D quadrilateral cell

6.3.3 Jacobian ratio

The Jacobian matrix describes the properties of the mapping between the computational space

(𝜉1, 𝜉2, 𝜉3), where the NS equations are discretized and solved, and the real domain (𝑋1, 𝑋2, 𝑋3).

In an ideal situation the two domains would coincide, thus the computed solution could be

transferred to the real case without loss of accuracy. For each element of the mesh the

determinant of the Jacobian matrix is computed at some sampling points (for example the corner

nodes, the centroid...).

Figure 34: mapping of an hexahedral element (Bucki, [5])

Page 51: Numerical Modelling of a Radial Inflow Turbine with and

40

From eqn.(5.6) JR is defined as the maximum to the minimum value among those determinants.

Other definitions presented in literature [5] consider the maximum determinant at denominator,

but here is reported the one used by ANSYS® Meshing.

𝐽𝑅 ≝𝑚𝑎𝑥 𝑑𝑒𝑡 𝐽 𝜉𝑖

𝑚𝑖𝑛 𝑑𝑒𝑡 𝐽 𝜉𝑖 (5.6)

where 𝜉𝑖 , 𝑖 ∈ 1,2,… ,𝑛 is the generic sampling point.

At the end JR is a measure of the maximum distorsion of each element of the mesh. The situation

is sketched if Fig.34. A value close to 1 indicates that the mapping does not lead to distorsion of

the elements, while the higher the Jacobian ratio the worse is the mesh.

6.4 Meshing of components

Volute, nozzle ring and diffuser were meshed using the software 𝐴𝑁𝑆𝑌𝑆®Meshing. The process

of mesh generation is highly automated: the designer specifies general sizing parameters such as

degree of fineness of the grid (relevance), rate at which adjacent elements are allowed to grow

(transition), control over the element quality (smoothing), and the software creates an

unstructured mesh based on those requirements and on constraints about the dimension of the

elements to be used (limits on the size of edges and faces), which are specified as defaults

(defaults may be changed if necessary). Fig.35 shows as an example the setup of

𝐴𝑁𝑆𝑌𝑆®Meshing for mesh generation on the volute.

Figure 35: example of setup of 𝐴𝑁𝑆𝑌𝑆®Meshing (volute)

Generation of unstructured mesh by means of 𝐴𝑁𝑆𝑌𝑆®Meshing is a fast and robust process,

however for a more accurate description of the flow in the boundary layer a structured mesh

close to the walls is needed. This is achieved in 𝐴𝑁𝑆𝑌𝑆®Meshing by using inflation layers: to

this purpose the program requires definition of the surface around which inflation must be

Page 52: Numerical Modelling of a Radial Inflow Turbine with and

41

performed (named selection) and specifications about inflation option (thickness of the first

layer, total inflation thickness…), number of layers and growth rate between adjacent layers. An

example is reported in Fig.35.

The main parameter to be set at this stage is the first layer thickness, i.e. the distance of the first

node of the mesh from the wall expressed in terms of 𝑦+. A target 𝑦+ is set and the

correspondent 𝑦 is determined through eqn.(5.2), but in order to do so the friction velocity 𝑢∗ must be estimated. Literature (White, [30]) reports a graph, namely Moody’s diagram, which

relates 𝑅𝑒 to 𝐶𝑓 (skin friction coefficient) for circular pipes with smooth walls in turbulent

regime. Many empirical relations were also formulated, among which the 1/7th

law is mentioned.

Assuming the validity of such relation in more general cases of ducts with non-circular shapes

(volute, diffuser…), once 𝑅𝑒 is calculated from eqn.(5.1) it is possible to estimate 𝐶𝑓 from

eqn.(5.7)

𝐶𝑓 = 0.027𝑅𝑒𝑥−1

7 (5.7)

and 𝑢∗ from eqn.(5.8)

𝑢∗ ≝ 𝜏0

𝜌= 𝐶𝑓

𝑈𝑟𝑒𝑓2

2 (5.8)

In Tab.6 are reported the target 𝑦+ for each sub-domain of the turbine (except for the rotor,

which will be discussed later) and the correspondent 𝑦, which must be set as first layer thickness

in the setup for generating inflation layers with 𝐴𝑁𝑆𝑌𝑆®Meshing. The reference values for the

computation of 𝑅𝑒 in eqn.(5.1) are considered to vary among the components.

Domain 𝑈𝑟𝑒𝑓 [𝑚/𝑠] 𝐿𝑟𝑒𝑓 [𝑚] 𝑅𝑒 [−] 𝑦+ [−] 𝑦 [𝑚]

Volute 𝑈1 ∼ 120 (speed at inlet)

3 ∗ 10−2 (radius at inlet)

4 ∗ 104 30 4 ∗ 10−4

Nozzle ring 𝑈𝑛𝑟−𝑖𝑛𝑙𝑒𝑡 ∼ 250 (speed at inlet)

3.2 ∗ 10−2 (blade chord)

8.8 ∗ 104 10 6 ∗ 10−5

Diffuser 𝑈3 ∼ 210 (speed at inlet)

2.1 ∗ 10−2 (width at inlet)

4.9 ∗ 104 30 2 ∗ 10−4

Table 6: estimation of first layer thickness for turbine components

Notice that for volute and diffuser 𝑦+ is located in the logarithmic layer while for the nozzle ring

it is placed in the buffer layer (𝑦+ ∼ 10). The goal is to achieve higher resolution in the nozzle

ring where the blades may cause separation of the flow, especially when the turbine is working at

off-design points.

DOMAIN NUMBER

OF NODES

NUMBER OF

ELEMENTS

AVERAGE

SKEWNESS

AVERAGE

JR

AVERAGE

OQ

Volute-

vaneless

113587 338479 0.256 1.069 0.872

Volute-

vaned

111830 336893 0.255 1.074 0.871

Nozzle ring 262591 995715 0.310 1.078 0.807

Diffuser 42571 122799 0.202 1.019 0.899

Table 7: mesh statistics for turbine components

Page 53: Numerical Modelling of a Radial Inflow Turbine with and

42

Once the inflation is defined the setup of 𝐴𝑁𝑆𝑌𝑆®Meshing is complete and meshes are

generated. The result is shown in Tab.7, where the statistics for the final meshes are reported. As

can be seen all meshes fulfil the quality criteria, thus they will be used for the subsequent CFD

analysis.

The implemented solution is illustrated in the figures below, whose goals are to show the effect

of inflation layers (Fig.36), an ensemble mesh (Fig.37) and the presence of local refinements

around “critical points” (Fig.38, at LE and TE of the nozzle ring blade).

Figure 36: mesh of the volute (section)

Figure 37: mesh of the diffuser (ensemble)

Page 54: Numerical Modelling of a Radial Inflow Turbine with and

43

Figure 38: mesh of the nozzle ring (one blade)

For meshing the rotor TURBOGRID® was used, a dedicated software for rotating machines which

creates a high quality structured hexahedral mesh. The process of mesh generation proceeds

according to the following steps:

1. The geometry of the rotor is imported from a BLADEGEN® file.

2. Based on the geometry of the blade TURBOGRID® automatically chooses the most

suitable topology for the mesh using the function ATM optimized. Topology denotes the

pattern in which the region of space around the blade is divided. Topology is chosen

according to the blade profile, the local curvature, the shape of LE and TE (cut-off or

rounded) and other geometrical factors: in each block the elements of the mesh are

oriented along the local shape of the blade, and interfaces between blocks are “smooth”

so to avoid warped elements which lower the quality of the mesh.

The topology for the rotor is sketched in Fig.39.

Figure 39: topology for the rotor blade. The blade (blue) is surrounded by meshing blocks

Page 55: Numerical Modelling of a Radial Inflow Turbine with and

44

3. At this step TURBOGRID® requires to specify the fineness of the overall mesh, through the

global size factor, and the target 𝑦+. For the rotor 𝑦+ = 1 is set: the purpose is to directly

resolve the boundary layer (without the use of wall functions), as higher accuracy is

needed in the rotor, where the flow is fully 3D.

4. The final mesh is automatically generated. The final mesh for the rotor has 613788 nodes

and 579535 elements.

Fig.40 shows a portion of the mesh around the LE of the blade. Progressive refinement can be

appreciated close to the wall, due to the specification on 𝑦+, while the topology is seen from the

orientation of the hexahedral elements.

Figure 40: mesh of the rotor (portion)

After mesh generation TURBOGRID® provides tools for assessing the quality of the mesh. The

most significant ones are listed below:

Minimum/Maximum face angle. For each face of an element, the angle between the two

edges that touch a node is calculated. The smallest/largest angle of all faces is returned as

the value for the minimum/maximum face angle. This parameter evaluates how warped

the faces are with respect to the ideal angle (90°), therefore it can be considered a

measure of skewness.

Minimum volume. This is the minimum volume among all the cells of the grid. Its value

must be always positive in order avoid numerical errors.

Maximum element volume ratio. For each node the volume of all the cells touching that

node is computed and the ratio between the maximum and the minimum volume is

returned. This is a measure of the local expansion factor, and it should be low especially

in regions where high gradients are expected in the flow quantities (velocity,

temperature...).

Maximum edge length ratio. For each face of an element, the ratio between the longest

and the shortest edge is computed and the maximum value is returned. This parameter

measures the aspect ratio (see Fig.32).

Mesh statistics for the rotor are reported in Fig.41: after computation TURBOGRID® checks if

each quality parameter lays within the acceptable range, and returns a feedback.

Page 56: Numerical Modelling of a Radial Inflow Turbine with and

45

Figure 41: mesh statistics for the rotor

The overall mesh for the vaneless configuration contains 769946 nodes and 1040813 elements,

for the vaned configuration 1030780 nodes and 2034942 elements.

Page 57: Numerical Modelling of a Radial Inflow Turbine with and

46

7 – CFX SETUP

Once generated and meshed, all components of the turbine are assembled in CFX® − Pre.

However, before running a CFD simulation, it is necessary to specify the mathematical model

that must be solved together with boundary conditions to the problem, initial conditions and

interfaces between various components. Such topics are treated in this chapter.

7.1 Mathematical model for turbulence

A turbulent flow is characterized by swirling structures (eddies) which span a wide spectrum of

scales (Johansson & Wallin, [8], 2012). The dimension of the largest eddies is set by the

reference geometrical size of the problem, while the length scale of the smallest eddies is

supposed to only depend on dissipation rate and viscosity: this is, infact, the scale at which the

flow dissipates kinetic energy through viscous mechanisms. Within the two limits there exist a

range of eddies of intermediate dimensions which transfer kinetic energy from the biggest to the

smallest scales in a process called turbulence cascade.

The ratio between the largest and the smallest scales of turbulence depends on the Reynolds

number and can be estimated (for details see reference [8]) as 𝑅𝑒3/4

. Thus, a grid which aims to

describe all the details in a 3D flow domain would have a resolution of 𝑅𝑒9/4

, and considering

that for the present application the Reynolds number is at least in the order of 104 the mesh for

DNS would have 109 nodes and the computational effort would be massive. Alternatively it is

possible to describe a turbulent flow by means of RANS, where it is assumed that turbulence can

be modelled as fluctuations within an average velocity field: however when this assumption is

introduced in the NS equations it originates an extra term, namely the Reynolds stress, which is

unknown hence it must be modelled through a turbulence model.

Common turbulence models are the so called 𝑘 − 휀 and 𝑘 − 𝜔: both solve 2 additional

equations, together with the set of RANS, which account for the transport of turbulent variables.

In the former case those variables are turbulent kinetic energy, k, (representing the variance of

the fluctuations in velocity) and turbulence dissipation, ε, (representing the rate at which velocity

fluctuations dissipate); in the latter case ε is replaced by the specific turbulence dissipation, ω,

(𝜔 ≝휀

𝑘). Both models rely on the assumption that the Reynolds stress term is related to the

gradient of the mean velocity (strain) through the turbulent viscosity, 𝜈𝑡 , according to the

Boussinesq hypothesis (notice that this hypothesis is purely empirical)

−𝑢𝑖′𝑢𝑗′ 𝑅𝑒𝑦𝑛𝑜𝑙𝑑𝑠 𝑠𝑡𝑟𝑒𝑠𝑠

= 𝜈𝑡 𝑡𝑢𝑟𝑏𝑢𝑙𝑒𝑛𝑡𝑣𝑖𝑠𝑐𝑜𝑠𝑖𝑡𝑦

∗ 𝑓 𝜕𝑈𝑖𝜕𝑥𝑗

𝑠𝑡𝑟𝑎𝑖𝑛

, 𝑖, 𝑗 ∈ 1,2,3 (6.1)

An in-depth description of 𝑘 − 휀 and 𝑘 − 𝜔 turbulence models is beyond the scope of this work,

however from the point of view of applications 𝑘 − 휀 is reported to have general good

performances but is not applicable to flows under adverse pressure gradients and is not accurate

in forecasting separation, while 𝑘 − 𝜔 has complementary characteristics and is mostly suitable

in the near-wall region. In this thesis a hybrid turbulence model, namely SST 𝑘 − 𝜔, is used, as it

combines the advantages of both models because it allows a shift from 𝑘 − 𝜔 to 𝑘 − 휀 depending

on the distance from the wall.

The SST 𝑘 − 𝜔 formulation is shown below:

Page 58: Numerical Modelling of a Radial Inflow Turbine with and

47

k-equation: 𝜕 𝜌𝑘

𝜕𝑡+

𝜕 𝜌𝑈𝑗𝑘

𝜕𝑥𝑗=

𝜕

𝜕𝑥𝑗 𝜇 +

𝜇 𝑡

𝜍𝑘3 𝜕𝑘

𝜕𝑥𝑗 + 𝑃𝑘 − 𝛽

∗𝜌𝑘𝜔 (6.2)

ω-equation:

𝜕 𝜌𝜔

𝜕𝑡+

𝜕 𝜌𝑈𝑗𝜔

𝜕𝑥𝑗=

𝜕

𝜕𝑥𝑗 𝜇 +

𝜇 𝑡

𝜍𝜔3 𝜕𝜔

𝜕𝑥𝑗 + 1− 𝐹1 2𝜌

1

𝜍𝜔2𝜔

𝜕𝑘

𝜕𝑥𝑗

𝜕𝜔

𝜕𝑥𝑗+ 𝛼3

𝜔

𝑘𝑃𝑘 − 𝛽3𝜌𝜔

2 6.3

where 𝑃𝑘 is the term responsible for the production of turbulence and depends on the Reynolds

stress (which is modelled as function of 𝜈𝑡).

𝐹1 ∈ [0,1] is called blending function and accounts for the position with respect to the wall. By

tuning 𝐹1 the model shifts from the expression of the standard 𝑘 − 휀 (when 𝐹1 = 0) to the one of

𝑘 − 𝜔 (when 𝐹1 = 1). The constants of the model (𝛽∗, 𝛽3, 𝜍𝑘3, ...) are a linear combination of

the coefficients appearing in the formulations of 𝑘 − 휀 and 𝑘 − 𝜔 through the blending function,

i.e. taken ϕ as a general constant it holds 𝜙𝑆𝑆𝑇 = 𝐹1𝜙𝑘−𝜔 + (1− 𝐹1)𝜙𝑘−휀 . The blending function is formulated empirically.

7.2 Near-wall treatment

In the previous chapter it was seen that, depending on how close the first node of the mesh is

from the wall, the flow in the boundary layer can be either resolved or modelled through wall

functions, and grids for all components of the turbine were designed based on the target 𝑦+.

However an exact estimation of 𝑦+ is difficult to achieve because 𝑦+ depends on 𝑅𝑒 , which

varies inside the turbine. In particular inaccuracies may occur if the wall function method is used

in regions where the first node of the mesh is inside the viscous sublayer, while if the Low

Reynolds Number method were used with the first node of the grid being in the log-law region it

would not be possible at all to describe viscous and buffer layer.

If the wall function is set to automatic, the correct near-wall treatment is automatically chosen by

CFX®. The program calculates 𝑦+ and if the mesh has a local near-wall distance corresponding

to 𝑦+ < 11.06 (default value, [22]) the boundary layer is resolved, otherwise wall functions are

used. In this way automatic wall function allows the highest possible accuracy.

7.3 Boundary conditions and interfaces

Boundary conditions are relative to the design point and are common to both vaneless and vaned

configurations. Boundary conditions are derived from the meanline design and reported below

Inlet total pressure: 2.5636 [bar]

Inlet total temperature: 845 [K]

Outlet static pressure: 1.1277 [bar]

Flow direction at inlet: normal to inlet section

Other combinations of boundary conditions are possible in CFX® (mass flow rate, velocity...) but

their use turned out to give solutions non-monotonically convergent (the residuals of mass and

momentum showed undamped oscillations) or even non convergent at all, while with the

specification of total pressure at inlet and static pressure at outlet the solution was stable. This

choice is also supported by literature ([9], [22])

No slip condition was set at the walls. Walls are modelled as adiabatic and smooth.

Page 59: Numerical Modelling of a Radial Inflow Turbine with and

48

Interfaces are used to model the contact region between components of the turbine: across them

there can be a discontinuity in the mesh pattern, a variation of frame of reference (for example in

the matching rotor-stator) or a variation of pitch (this is the case when two contact domains have

different angular width).

For the present study the stage model provided by CFX® is used. In the stage model, also called

mixing plane model, the flow properties (i.e. velocity, pressure, temperature...) are averaged

circumferentially upstream of the interface in order to obtain the boundary condition for the

downstream component. Since the flow is averaged, the inlet boundary condition for the

component downstream of the interface is steady, thus this method is suitable for steady state

simulations, which is the case here.

Another type of interface is the rotational periodicity in a rotor SBP, which allows to model the

whole rotor as just one blade passage, thus saving computational time. All interfaces and

boundary conditions are highlighted in the ensemble Fig.42 (only the vaned configuration is

illustrated, as it is the most complete case).

Figure 42: illustration of interfaces and boundary conditions for vaned configuration

7.4 Choice of off-design points

The design point represents average conditions of inlet pressure and temperature, but within the

combustion cycle these parameters vary over a considerable range, as seen from Fig.43.

Off-design points were determined [14] by measuring the flow properties of the exhausts at

constant intervals of 2.32 milliseconds after the opening of the discharge valve: among them

only the most representative are chosen for this study, and are listed in Tab.8

Page 60: Numerical Modelling of a Radial Inflow Turbine with and

49

Figure 43: choice of representative off-design points (Mora, [14], adapted)

In particular points 1, 4, 5 are associated with the “furthest” thermodynamic conditions with

respect to the average, while point 7 is the closest to it and is chosen to represent the effect that

small variations around the mean inlet conditions may have on the performance of the turbine.

Notice that the highest temperature of the combustion cycle is not associated with the highest

pressure, hence both points are studied.

POINT TEMPERATURE (total) [K] PRESSURE (total) [bar]

1 762.9 2.0754

4 993.1 2.8438

5 935.1 3.2098

7 829.1 2.6339

Table 8: thermodynamic properties of the studied off-design points

Page 61: Numerical Modelling of a Radial Inflow Turbine with and

50

8 – RESULTS

This section presents the main results of the steady simulations run with CFX® on the designed

CAD model. For each working condition (design and off-design) the turbine is analysed in both

configurations, with and without nozzle ring, and results are compared. Such results will then be

discussed in Chapter 8 in order to answer to the stated objective.

8.1 Design point

PARAMETER VANELESS VANED

Mass flow [kg/s] 0.396 0.417

Torque on rotor [Nm] (*) -5.11 -5.29

Power [kW] 45.47 47.07

Efficiency T-T [-] 0.751 0.762

Table 9: performance comparison, design point

(*) The negative sign is a convention, as the sense of rotation of the impeller is clockwise.

Performance data relative to the design condition are calculated and reported in Tab.9, while

Tab.10 illustrates the mean velocity triangles at rotor inlet and outlet.

PARAMETER

(rotor inlet,

average)

VANELESS

(red)

VANED

(blue)

(*)

𝑈2 [𝑚/𝑠] 365 358

𝑐2 [𝑚/𝑠] 333 324

𝑤2 [𝑚/𝑠] 176 161

𝛼2 [°] 61.3 63.2

𝛽2 [°] -24.6 -24.7

PARAMETER

(rotor outlet,

average)

VANELESS

(red)

VANED

(blue)

𝑈3 [𝑚/𝑠] 153 161

𝑤3 [𝑚/𝑠] 398 403

𝑐3 [𝑚/𝑠] 311 311

𝛽3 [°] -38.8 -39.9

𝛼3 [°] -15.8 -15.96

Table 10: comparison between mean velocity triangles at rotor inlet (top) and at rotor outlet

(bottom), design point

(*) Sketches represent the projection in the meridional plane (where α and β are defined) but c and w are general

3D vectors and also have an out-of-plane component.

Page 62: Numerical Modelling of a Radial Inflow Turbine with and

51

The spanwise distribution of 𝛽2 is illustrated in Fig.44, while distribution of 𝛼3 in Fig.45. For

each spanwise position the correspondent value is obtained as a mass average over the

circumferential coordinate.

Figure 44: spanwise distribution of 𝛽2, comparison at design point

Figure 45: spanwise distribution of 𝛼3, comparison at design point

Page 63: Numerical Modelling of a Radial Inflow Turbine with and

52

Figure 46: rotor blade loading, comparison at 10% span (top, left), 50% span (top, right), 90%

span (bottom), design point

Figure 47: static entropy around the rotor blade at 90% span. Left: vaneless. Right: vaned

Page 64: Numerical Modelling of a Radial Inflow Turbine with and

53

Figure 48: Mach distribution around the rotor blade at 10% span. Top: vaneless. Bottom: vaned

Figure 49: Mach distribution around the rotor blade at 50% span. Left: vaneless. Right: vaned

Figure 50: Mach distribution around the rotor blade at 90% span. Left: vaneless. Right: vaned

Page 65: Numerical Modelling of a Radial Inflow Turbine with and

54

Fig.46 shows the comparison of blade loading for three representative spanwise positions: close

to hub (10%), mid span (50%) and close to shroud (90%). Fig.48-49-50 allow to visualize the

blade-to-blade velocity distributions (in terms of Mach number) at the same spanwise locations

at which the blade loading was evaluated. Fig.47, instead, is useful to visualize the static entropy

at a particular section in which flow separation is detected.

Fig.51 shows the flow around a blade of the nozzle ring at design point. The blade in the picture

is the closest to the volute tongue.

Figure 51 : velocity contour around nozzle ring blade, 50% span, design point

8.2 Off-design points

POINT CONFIGURATION Mass flow

[kg/s]

Torque on

rotor [Nm]

Power

[kW]

Efficiency T-T

[-]

1 vaneless 0.312 -2.88 25.64 0.757

vaned 0.321 -2.93 26.06 0.769

4 vaneless 0.428 -6.56 58.43 0.707

vaned 0.438 -6.86 61.10 0.729

5 vaneless 0.500 -8.08 71.89 0.795

vaned 0.506 -8.03 71.46 0.712

7 vaneless 0.424 -5.51 49.03 0.825

vaned 0.430 -5.51 49.03 0.763

POINT CONFIGURATION 𝛽2 [°] 𝛼3 [°]

1 vaneless -41.85 -15.68

vaned -44.32 -12.14

4 vaneless -13.25 -20.96

vaned -10.29 -21.89

5 vaneless -13.78 -22.38

vaned -14.04 -22.37

7 vaneless -22.28 -18.01

vaned -25.51 -16.42

Table 11: comparison of performance (top) and mean flow angles (bottom) at off-design points

Page 66: Numerical Modelling of a Radial Inflow Turbine with and

55

8.2.1 Off-design point 1

Figure 52: spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 1

Figure 53 : Velocity distribution around the rotor blade at 50% span. Left: vaneless. Right:

vaned

Figure 54: Velocity distribution around the rotor blade at 90% span. Left: vaneless. Right:

vaned

Page 67: Numerical Modelling of a Radial Inflow Turbine with and

56

8.2.2 Off-design point 4

Figure 55: spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 4

Figure 56: static entropy in meridional plane, circumferential average, off-design point 4. Left:

vaneless. Right: vaned

Figure 57: Velocity distribution around the blade (left) and blade loading (right) at 90% span,

off-design point 4

Page 68: Numerical Modelling of a Radial Inflow Turbine with and

57

8.2.3 Off-design point 5

Figure 58: spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 5

Figure 59: static entropy in meridional plane, circumferential average, off-design point 5. Left:

vaneless. Right: vaned

Figure 60: velocity contour around nozzle ring blade, 50% span, off-design point 5

Page 69: Numerical Modelling of a Radial Inflow Turbine with and

58

8.2.4 Off-design point 7

Figure 61: spanwise distribution of 𝛽2 (left) and 𝛼3 (right), off-design point 7

Figure 62: rotor blade loading, comparison at 10% span (top, left), 50% span (top, right), 90%

span (bottom), off-design point 7

Page 70: Numerical Modelling of a Radial Inflow Turbine with and

59

9 – DISCUSSION

In this section results are analysed: the goal is to investigate the effects of the static nozzle ring

on the performance of the radial turbine. In order to do a “fair” comparison, at design point both

vaned and vaneless configurations must have the same average inflow conditions at rotor inlet.

Mean velocity triangles are reported in Tab.10. It holds 𝑐2_𝑣𝑎𝑛𝑒𝑑

𝑐2_𝑣𝑎𝑛𝑒𝑙𝑒𝑠𝑠= 0.973,

𝛽2_𝑣𝑎𝑛𝑒𝑑

𝛽2_𝑣𝑎𝑛𝑒𝑙𝑒𝑠𝑠= 1.004,

hence between the two cases the absolute velocity varies by less than 3% and the relative inflow

angle by less than 1%. The condition is fulfilled. Similar situation is present at rotor outlet,

where again the mean velocity triangles show non-appreciable differences (Tab.10, bottom).

According to a meanline analysis there would be no difference in incidence losses or TE losses

between vaned and vaneless configurations.

Fig.44 shows the spanwise distribution of 𝛽2. In both cases the relative inflow angle drops near

hub and shroud, while it is above the average value around mid span. At each spanwise position,

a strong correlation is noticed between the local 𝛽2 and the flow velocity around the blade: this

was an expected behaviour, as the inflow angle affects the position of the stagnation point at the

LE of the local blade profile and consequently the whole pressure distribution. At 90% span it

holds 𝛽2_𝑣𝑎𝑛𝑒𝑑 > 𝛽2_𝑣𝑎𝑛𝑒𝑙𝑒𝑠𝑠 (Fig.44): in the vaned turbine the stagnation point is shifted towards

the pressure side of the blade (the flow is more radial, which represents the ideal nominal design

condition), and as can be seen from Fig.50 this causes the flow on the suction side to separate

closer to the TE. This is in accordance with the correspondent blade loading plot (Fig.46,

bottom), where pressure on the suction side keep decreasing in the vaned case (blue curve) while

in the vaneless case the blade profile has already stalled (red curve). From Fig.46 it can be seen

that the rotor blade is more loaded in the vaned turbine: as blade loading determines the moment

on the impeller about the axis of rotation, the higher blade loading justifies the higher power

output of the vaned case (Tab.9). The difference is about 3.5%. Separation not only affects the

blade loading but also the efficiency. Regions of separated flow are turbulent and strong

dissipation of kinetic energy occurs. This is visualized in Fig.47, which represents the entropy of

the flow in the blade-to-blade plane (90% span). Higher entropy in the vaneless case denotes

higher amount of losses with respect to the turbine with nozzle ring: this difference is indeed

confirmed by the performance calculations and is about 1.5% (Tab.9).

In the vaned configuration separation may occur at nozzle ring blades. This is not the case at

design point, except for the blade downstream of the volute tongue. Here the free-vortex law,

which was the underlying assumption for the design of the volute, is not valid [32] and the flow

angle at volute outlet is higher than the design value. However the region subject to separation is

very small (see Fig.51).

At off-design points both configurations undergo substantial modifications of average inflow and

outflow angles with respect to the design value (Tab.11, bottom). This first observation suggests

that the nozzle ring does not constrain 𝛽2 to the design value (or close values) also at off-design

points. Inspection of Fig.52, 55, 58, 61 shows that not only the mean value but also the spanwise

distribution of 𝛽2 (and 𝛼3) vary considerably. Another straightforward consideration comes from

Tab.11: it does not exist a turbine configuration which guarantees better performance (in terms

of efficiency and power output) at all working conditions.

Efficiency seems to be related to the spanwise distribution of 𝛽2. Let’s consider off-design point

4. The blue curve (vaned) is always above the red one (vaneless), which means that 𝛽2 is closer

to 0 (ideal case) through all the span. A plot in the meridional plane (Fig.56) shows that this

condition corresponds to lower amount of static entropy, especially close to hub and shroud

(where, according to Fig.55, left, the difference in inflow angle increases even more). As a

Page 71: Numerical Modelling of a Radial Inflow Turbine with and

60

consequence the efficiency of the vaned turbine is higher than vaneless for this off-design point

(around 3.1%). The relashionship between rotor inflow angle and entropy should be investigated

in details, but the intuitive explanation is that the blade has been designed under the condition

𝛽2,𝑏 = 𝛽2 = 0 at all spanwise locations and the less this condition is satisfied the more the flow

is subject to detachment or recirculation, which leads to an increase of entropy. The situation is

visualized in Fig.57: close to the shroud (90% span) both configurations separate at TE (the

velocity around the blade drops at streamwise position near 0.7) but the vaned separates after,

hence not only flow entropy is lower (Fig.56, shroud) but also blade loading is higher (Fig.57,

right) and ultimately power output is higher.

The same argumentation can be followed for off-design point 5, but here 𝛽2_𝑣𝑎𝑛𝑒𝑙𝑒𝑠𝑠 >𝛽2_𝑣𝑎𝑛𝑒𝑑 everywhere except close to hub and shroud (Fig.58, left). The meridional plot of entropy

shows the forecasted behaviour (Fig.59) and efficiency is higher for the vaneless configuration

(Tab.11). Moreover in this case the region of separated flow at the blade of the nozzle ring is

bigger than at design point (again the difference is more evident for the blade close to volute

tongue).

At off-design point 1 the inflow angle is large and separation occurs both at LE and TE (Fig.54).

The situation regarding 𝛽2 and consequent flow pattern can be interpreted as the previous cases,

however here the difference in 𝛼3 between vaned and vaneless turbine is not negligible and as

𝛼3_𝑣𝑎𝑛𝑒𝑙𝑒𝑠𝑠 < 𝛼3_𝑣𝑎𝑛𝑒𝑑 (mean value) a difference in TE losses must be considered together with

losses due to 𝛽2. Overall efficiency for the vaned case is slightly higher (+1.6%) than vaneless.

Off-design point 7 falls inside the frame used to describe the previous cases.

Regarding mass flow at design point it can be noticed that the value for both turbine

configurations is higher with respect to the meanline value that was used during the design

phase. This difference has already been pointed out in literature [22]. The most plausible reason

is that the meanline code over-predicts the effects of blockage inside the rotor blade channel. As

can be seen from blade-to-blade Mach plots (Fig.48, 49, 50), Mach can be locally 1 (or higher),

especially near the shroud, but the channel is never chocked at all spanwise positions, not even

for off-design points implying the highest values of inlet pressure and temperature (in this case

points 4 and 5). The reason why mass flow in the vaned configuration is slightly higher than

vaneless at all working conditions may be linked to the inflow angle 𝛽2 near the hub: in the

vaneless case this stays well below the average (while in the vaned it tends to increase more

rapidly) and such low values of the inflow angle locally cause recirculation. Somehow chaotic

streamlines can be seen near the hub, and this may support the statement, but as this cannot be

proved a deeper investigation should be done on the topic.

This analysis shows that the reason why a turbine configuration is “better” than another one is

mainly related to the inflow angle at rotor inlet, which should be somehow close to the design

value for the rotor blade (in this case 0°). If the presence of a static nozzle ring guarantees this

condition in a certain working point the vaned configuration is preferable to the vaneless,

otherwise not, but from this analysis it is possible to conclude that the addition of a static nozzle

ring in a 90° IFR turbine does not improve performance in all the combustion cycle.

Page 72: Numerical Modelling of a Radial Inflow Turbine with and

61

10 – CONCLUSION AND FUTURE WORKS

In this thesis a comparison between vaneless and vaned turbine configurations has been made on

the basis of numerical simulations carried out in steady conditions with a CFD commercial

software. The use of a static nozzle ring increases both efficiency and power output at design

point (increment is roughly estimated in +1.5% and +3.5% respectively). On the other hand, off-

design performances strongly depend on the thermodynamic flow conditions which characterize

the specific point, and a general trend does not exist: however it is proved that the

implementation of a fixed nozzle ring in a radial turbine does not guarantee higher efficiency

through all the engine combustion cycle.

Inputs for possible future works are listed below:

Results presented in this thesis solely rely on a numerical model, whose uncertainties

have been highlighted in Chapter 4 – Limitations. In order to assess the validity of the

results the latter should be compared with experimental data

In this study only steady conditions have been considered, and it is implicitly assumed

that the flow has enough time to adapt to variations of pressure and temperature in the

exhaust gases. This may not be true in general, especially when the engine operates at

high rpm, hence a further study should model the unsteadiness of the flow

Future efforts could be made for improving the 3D geometrical model of the radial

turbine: for example the casing which surrounds the rotor is not modelled here, which

implies a poor evaluation of tip clearance losses

No structural considerations are made regarding the design of the turbine. The rotor, in

particular, is subject to high centrifugal forces, and despite the blade tip speed is limited

to 𝑈2 < 400 𝑚/𝑠 under design parameters (𝑈2 ~ 356 𝑚/𝑠), there is no guarantee that

the blade thickness distribution suggested in this thesis results in mechanical stresses

which do not lead to fracture or plastic deformation of the blade. For this reason a further

study is needed in order to check the feasibility of the present design.

Page 73: Numerical Modelling of a Radial Inflow Turbine with and

62

11 – BIBLIOGRAPHY

[1] M. Abidat, M. K. Hamidou, M. Hachemi, M. Hamel, “Design and Flow Analysis of Radial

and Mixed Flow Turbine Volutes”, European Conference on Computational Fluid

Dynamics, TU Delft, The Netherlands, 2006

[2] N. C. Baines, M. Lavy, “Flow in Vaned and Vaneless Stators of Radial Inflow

Turbocharger Turbines”, Proceedings Institution Mechanical Engineers, Turbochargers and

Turbocharging Conference, 1969

[3] A. Bakker, “Course Material and Lectures”, Dartmouth, 2002-2006, available at

http://www.bakker.org

[4] R. D. Blevins, “Applied Fluid Dynamics Handbook”, Krieger Publishing Company, 1984

[5] M. Bucki, C. Lobos, Y. Payan, N. Hitschfeld, “Jacobian-Based Repair Method

for Finite Element Meshes after Registration”, Engineering with Computers, Springer

Verlag, pp.285-297, 2011

[6] S. L. Dixon, C. A. Hall, “Fluid Mechanics and Thermodynamics of Turbomachinery”, 6th

Edition, Butterworth-Heinemann, 2010

[7] M. S. Floater, “Beziér Curves and Surfaces”, Lecture Notes, Oslo, 2003

[8] A. V. Johansson, S. Wallin, “Turbulence Lecture Notes”, Stockholm, 2012

[9] M. Khader, “Optimized Radial Turbine Design D1.8”, School of Mathematics, Computer

Science and Engineering, London, 2014

[10] A. Khare, A. Singh, K. Nokam, “Best Practice in Grid Generation for CFD Applications

Using HyperMesh”, available at http://www.altairatc.com

[11] P. K. Kundu, I. M. Cohen, “Fluid Mechanics”, 2nd

Edition, Academic Press, 2002

[12] S. A. MacGregor, A. Whitfield, A. B. Mohd Noor, “Design and Performance of Vaneless

Volutes for Radial Inflow Turbines. Part 3: Experimental Investigation of the Internal Flow

Structure”, Proceedings of the Institution of Mechanical Engineers Part A Journal of

Power and Energy, June 1994

[13] C. A. de Miranda Ventura, “Aerodynamic Design and Performance Estimation of Radial

Inflow Turbines for Renewable Power Generation Applications”, PhD Thesis, University

of Queensland, School of Mechanical and Mining Engineering, 2012

[14] E. C. Mora, “Variable Stator Nozzle Angle Control in a Turbocharger Inlet”, MSc Thesis,

KTH School of Industrial Engineering and Management Energy Technology, 2015

[15] G. Negri di Montenegro, M. Bianchi, A. Peretto, “Sistemi Energetici e Macchine a

Fluido”, Pitagora Editrice, Bologna, 2009

[16] H. Nguyen-Schäfer, “Rotordynamics of Automotive Turbochargers”, Springer Berlin

Heidelberg, 2012

Page 74: Numerical Modelling of a Radial Inflow Turbine with and

63

[17] M. H. Padzillah, M. Yang, W. Zhuge, R. F. Martinez-Botas, “Numerical and Experimental

Investigation of Pulsating Flow Effect on a Nozzled and Nozzleless Mixed Flow Turbine

for an Automotive Turbocharger”, ASME Turbo Expo 2014: Turbine Technical

Conference and Exposition, 2014

[18] S. Rajoo, R. Martinez-Botas, “Variable Geometry Mixed Flow Turbine for Turbochargers:

An Experimental Study”, International Journal of Fluid Machinery and Systems, Vol.1,

No.1, 2008

[19] H. I. H. Saravanamuttoo, H. Cohen, G. F. C. Rogers, “Gas Turbine Theory”, 4th

Edition,

Longman Group Limited, 1996

[20] S. P. Shah, S. A. Channiwala, D. B. Kulshreshtha, G. Chaudhari, “Design and Numerical

Simulation of Radial Inflow Turbine Volute”, De Gruyter, 2014

[21] S. Shah, G. Chaudhri, D. Kulshreshtha, S. A. Channiwala, “Effect of Flow Coefficient and

Loading Coefficient on the Radial Inflow Turbine Impeller Geometry”, International

Journal of Research in Engineering and Technology, Vol. 02, Issue 02, 2013

[22] E. M. V. Siggeirsson, S. Gunnarsson, “Conceptual Design Tool for Radial Turbines”, MSc

Thesis, Chalmers University of Technology, Department of Applied Mechanics,

Gothenburg, 2015

[23] A. T. Simpson, “Aerodynamic Investigation of Different Stator Designs for a Radial

Inflow Turbine”, Queen’s University Belfast, 2007

[24] K. J. Smith, J. T. Hamrick, “A Rapid Approximate Method for the Design of Hub Shroud

Profile of Centrifugal Impellers of Given Blade Shape”, NASA Technical Note 3399, 1954

[25] S. W. T. Spence, R. S. E. Rosborough, D. W. Artt, G. Mccullough, “A Direct Performance

Comparison of Vaned and Vaneless Stators for Radial Turbines”, ASME Journal of

Turbomachinery, Vol. 129, 2007

[26] W. D. Tjokroaminata, “A Design Study of a Radial Inflow Turbines with Splitter Blades

in Three-Dimensional Flow”, MSc Thesis, Massachussets Institute of Technology, 1992

[27] W. F. Torrey, D. Murray, “An Analysis of the Operational Costs of Trucking: 2015

Update”, American Transportation Research Institute, 2015

[28] R.A. Van den Braembussche, “Optimization of Radial Impeller Geometry”, Educational

Notes RTO-EN-AVT-143, Paper 13, Neuilly-sur-Seine, 2006

[29] Z. Wei, “Meanline Analysis of Radial Inflow Turbines at Design and Off-design

Conditions”, MSc Thesis, Ottawa-Carleton Institute for Mechanical and Aerospace

Engineering, 2014

[30] F. M. White, “Viscous Fluid Flow”, 2th

Edition, Mc Graw Hill, 1991

[31] A. Whitfield, A. B. Mohd Noor, “Design and Performance of Vaneless Volutes for Radial

Inflow Turbines. Part 1: Non-Dimensional Conceptual Design Considerations”,

Proceedings of the Institution of Mechanical Engineers Part A Journal of Power and

Page 75: Numerical Modelling of a Radial Inflow Turbine with and

64

Energy, June 1994

[32] A. Whitfield, S. A. MacGregor, A. B. Mohd Noor, “Design and Performance of Vaneless

Volutes for Radial Inflow Turbines. Part 2: Experimental Investigation of the Mean Line

Performance – Assessment of Empirical Design Parameters”, Proceedings of the Institution

of Mechanical Engineers Part A Journal of Power and Energy, June 1994

[33] S. M. Yahya, “Turbines, Compressors and Fans”, 4th

Edition, Mc Graw-Hill Education,

New Dehli, 2011

[34] Y. L. Yang, “A Design Study of a Radial Inflow Turbines in Three-Dimensional Flow”,

PhD Thesis, Massachussets Institute of Technology, 1991

[35] S. Hosder, B. Grossman, R. T. Haftka, W. H. Mason, L. T. Watson, “Remarks on CFD

Simulation Uncertainties”, MAD Center Report 2003-02-01, Virginia Polytechnic Institute

and State University, Blacksburg, USA

Page 76: Numerical Modelling of a Radial Inflow Turbine with and

65

APPENDIX 1: ABSOLUTE ANGLE AT ROTOR INLET

Page 77: Numerical Modelling of a Radial Inflow Turbine with and

66

Figure 63: spanwise distribution of 𝛼2. From top to bottom: design point, off-design 1, off-

design 4, off-design 5, off-design 7

Notice that the spanwise distributions of α2 and β2 are strongly related, since at rotor inlet the

blade speed is constant through all the span (points are equidistant from the rotational axis);

however plots of α2 are reported as well, for completeness.