In order to realistically assess the turbine efficiency, it was necessary to take into account the actual operating conditions; most of the time, in effect, the turbine does not operate at the design condition, since the pressure ratio and/or the rotational speed may be substantially different from the design values. For a given turbine geometry, the off-design operation will cause different mass flow rates and velocity triangles with respect to the design condition, and as a consequence, the turbine will exhibit a different overall performance, i.e., different output power and efficiency with respect to the design point. With the aim to calculate the performance of the turbine for any possible operating condition, it was necessary to evaluate the energy losses connected to the flow inside the machine by identifying the entropy generation mechanisms and quantifying their contribution in every phase of the fluid flow through the machine. The major contributions to the entropy generation [
19,
21] are caused by incidence losses, passage losses (due to friction, secondary flow and flow separation), tip leakage flow, and trailing edge losses, which occur in the internal blade passages in each of the elements of the turbomachine (volute, nozzle and rotor). All of these losses are conveniently quantified in terms of differences between the actual static enthalpy at the exit of a stage and that of the equivalent ideal isentropic process.
2.1. Analysis of the Volute
The purpose of the volute is to distribute the flow around the periphery of the turbine providing uniform mass flow rate and uniform static pressure at the volute exit, thus ensuring the proper exploitation of each rotor blade passage and avoiding unsteady radial loading of the rotating system.
The static temperature at volute inlet
T1 can be calculated from the application of the first law of thermodynamics to a perfect gas:
where
c1 is the absolute gas velocity at the volute inlet Section 1, and
is the specific heat at constant pressure evaluated at the average temperature
= (
TT1 + T1)/2 through the use of the Shomate equation, and the coefficients available in the NIST Chemistry WebBook [
22]. It is worth noting that Equation (1) implies an iterative calculation, since
requires the value of
T1, which is the output of the equation. As a starting value,
was assumed. Analogous iterations have been carried out for similar evaluations performed in this paper.
Since for any given turbine pressure ratio βT = PT1/P5 and rotor speed n, the pressure levels in the Section 1, Section 2, Section 3 and Section 4 (i.e., P1, P2, P3, and P4) are not known a priori, all the absolute velocities in the same sections are initially unknown; for this reason, each calculation procedure was started assuming an initial first attempt value of 100 m/s for each of the unknown absolute velocities (c1, c2, c3, and c4); the values of these velocities were then updated by successive iterations performed with the aim to obtain the mass-flow convergence on all the turbine sections.
Assuming an isentropic expansion from the total condition upstream from the turbine (gas pressure
PT1 and temperature
TT1) to the static condition, it was possible to calculate the static pressure at the volute inlet as:
where
is the mean isentropic coefficient evaluated at the average temperature
:
where, according to Meyer’s relation, the specific gas constant
R′ is the difference between the constant pressure and the constant volume specific heats
cp and
cv. It is worth pointing out that, in the case of the isentropic transformation, the average temperature
= (
TT1 + T1,is)/2 should be considered; however, since the difference between
T1 and
T1,is not relevant, it results that the difference between
and
is negligible (less than 0.01%), as is the difference between
and
; the same conclusion is obtained considering the turbine nozzle and the rotor, where the enthalpy drops are higher; according to this observation the authors adopted the approximation to use the average temperature of the actual evolution in place of the average temperature of the isentropic evolution when calculating the thermochemical properties of the gas, i.e., the specific heat at constant pressure and the isentropic coefficient.
The static density of the gas at the volute inlet
ρ1 was then calculated by means of the ideal gas law as function of
P1 and
T1 as:
Additionally, the mass-flow rate at the volute inlet
G1 is:
where
A1 is the inlet volute area. The absolute gas velocity
c1 was then corrected with the aim to reduce the difference between the mass flow at the volute inlet
G1 and outlet
G2 (whose calculation is shown below), which is considered a mass-flow error:
where
fC is the factor adopted for the absolute velocity iterative corrections; the same approach was followed for the correction of all the other absolute velocities, each one correlated to a proper mass-flow error; once all the mass-flow errors reach a negligible value (i.e., less than 0.1% of the mass-flow), the solution is considered the final and the calculation procedure is stopped. The convergence on the mass flow was reached acting on the absolute velocities (
c1,
c2,
c3, and
c4) rather than on the static pressure levels (
P1,
P2,
P3, and
P4) since the iterative correction performed on the static pressure values revealed some calculation instabilities.
Concerning the flow within the turbine volute, the most common assumption usually made is to consider a free vortex evolution; various extensions to the free vortex model, however, have been proposed in the scientific literature to reproduce the actual flow conditions, which are affected by friction, secondary flows, and mixing of the mainstream flow with outer fluid in the proximity of the outlet section. Loss correlations and coefficients may be used to account for deviations from the ideal evolution. Kastner and Bhinder [
23] represented the volute loss as a friction loss using conventional pipe-flow correlations. As proposed by Baines [
21], this loss item can be considered as the difference between the actual static enthalpy at exit from the volute and that of the equivalent isentropic process:
where
is the mean passage velocity (average value between inlet and outlet velocity),
Lhyd,vol and
Dhyd,vol are the hydraulic length and diameter of the volute respectively, while the coefficient of friction
Cf is defined as:
where
is the mean kinematic viscosity of the fluid between the inlet and outlet of the volute. The free vortex evolution can be hence modified to include the effects of shear stresses due to friction by means of the following relation:
which gives the tangential component of the absolute velocity at the volute outlet
c2u. This correlation is based on an analysis performed by Stanitz [
24] on the flow in the vaneless diffuser of a compressor, but being an analysis based on fundamental control volume, it could be adapted to the flows in the turbine volutes. Obviously, the volute loss produces a reduction in the actual fluid velocity at the volute outlet with respect to the ideal evolution, as shown by the simple application of the first law of thermodynamics between volute inlet and outlet sections:
where Δ
His(vol) and Δ
Hre(vol) are the isentropic and the actual enthalpy drop in the volute, respectively. As already mentioned, however, the pressure level at the volute outlet
P2 is initially unknown, which means that the isentropic enthalpy drop is also unknown and therefore the absolute velocity
c2 as well, which, in turn, is required for the calculation of the volute loss. The calculation procedure is hence based on successive iterative approximation; the initial value of 100 m/s was assumed for the absolute velocity at nozzle inlet
c2, which, in turn, allowed for evaluating the gas temperature at the volute outlet
T2 by the application of the first law of thermodynamics:
and the consequent actual enthalpy drop in the volute Δ
Hre(vol):
Considering the isentropic transformation, the static pressure at nozzle outlet was evaluated as:
where the isentropic enthalpy drop is obtained as the sum of the actual enthalpy drop (Equation (12)) and the volute loss (Equation (7)):
As already discussed, the mean constant pressure specific heat
and the mean isentropic coefficient
were both evaluated at the average temperature
= (
T1 + T2)/2. As done for the volute inlet (see Equation (4)), the fluid density at the volute outlet section
ρ2 was evaluated though the application of the ideal gas law as a function of the pressure
P2 and the temperature
T2. The mass flow through the volute outlet section
G2 could be hence calculated as:
where the meridional component of absolute velocity at volute outlet c
2m was evaluated as a function of the tangential component of the absolute velocity
c2u obtained by Equation (9):
Again, the absolute velocity at the volute outlet c
2 was corrected with the aim to reduce the mass-flow error between the volute outlet and nozzle outlet, i.e., (
G2 −
G3):
Once the value of the absolute velocity at volute outlet c2 is updated, then the entire calculation procedure from Equation (7) to Equation (17) is repeated until the mass-flow error (G2 − G3) becomes negligible (i.e., less than 0.1% of G2).
2.2. Analysis of the Nozzle
The nozzle consists of an annular ring of vanes that set the angle of approach of the working fluid to the rotor and it works in conjunction with the volute to accelerate the fluid. As in the case of the volute, nozzle losses ultimately lead to a reduction in outlet velocity compared to the ideal evolution. It is possible to distinguish nozzle passage losses and nozzle incidence losses. The nozzle passage loss is mainly caused by the action of friction forces between the flow and the nozzle blade’s solid surface [
25], which was determined by the following equation:
where
Lhyd,nozzle and
Dhyd,nozzle are the hydraulic length and diameter of the nozzle, respectively,
is the mean passage velocity (average value between nozzle inlet and outlet), and
f is the friction factor, evaluated as:
where
is the average Reynolds number between the nozzle inlet and outlet, while
RR represents the wall relative roughness, for which the value of 0.0002 m is suggested by Suhrmann et al. [
26].
The nozzle incidence loss, on the other hand, arises whenever there is a difference between the fluid-dynamic nozzle inlet angle α
2,f and the nozzle geometric inlet angle α
2,g; a simple yet effective correlation for the calculation of the nozzle incidence loss is:
As already observed in the case of the volute, the absolute velocity at nozzle outlet
c3 is related to the isentropic enthalpy drop in the nozzle Δ
His(nozzle), to the actual enthalpy drop Δ
Hre(nozzle) and to the losses Δ
hp(nozzle) and Δ
hi(nozzle) by the first law of thermodynamics applied between the inlet and outlet sections of the nozzle (i.e., Section 2 and Section 3 of
Figure 3):
Additionally, in this case, because the static pressure at the nozzle exit
P3 and the isentropic enthalpy drop are initially unknown, the absolute velocity at the nozzle outlet
c3 could not be evaluated through Equation (21); the calculation procedure was also based on a successive iterative approximation in this case, adopting a first attempt value of 100 m/s also for the absolute velocity
c3. The static temperature at nozzle outlet
T3 was hence calculated as:
and the related actual enthalpy drop in the nozzle:
The static pressure
P3 at nozzle outlet was obtained through the isentropic transformation:
where the isentropic enthalpy drop Δ
His(nozzle) was obtained from the actual enthalpy drop and the losses in the nozzle:
The static density of the fluid at nozzle outlet
ρ3 was evaluated by applying the ideal gas law (see Equation (4)) as a function of the temperature
T3 and pressure
P3, thus allowing evaluation of the mass flow at the nozzle exit
G3:
where
c3m is the meridional component of the gas velocity in the nozzle outlet section:
and
A3 the flow section normal to
c3m. As already carried out for the volute, the absolute velocity
c3 was corrected on the basis of the mass-flow error:
and hence the calculation procedure from Equation (18) to Equation (28) was repeated until the mass-flow convergence was reached.
2.3. Analysis of the Nozzle–Rotor Interspace
In a variable nozzle turbine, the swallowing capacity is controlled though the variation in the nozzle passage section, which is obtained by the rotation of the nozzle blades; the lower mass flows can be obtained adopting small opening angles; increasing the opening angles allows the turbine to swallow higher mass flows. For this reason, a radial gap Δ
r between the nozzle exit radius
r3 and the rotor inlet radius
r4 is necessary (see
Figure 4) to allow the nozzle blades to rotate without causing any impact on the rotor.
As can also be observed in
Figure 4, the radial gap increases when the nozzle blades reduce the opening angle, i.e., when the mass flow is reduced. This radial gap generates a limited interspace between the nozzle outlet (Section 3) and rotor inlet (Section 4), where it is plausible to assume an ideal free vortex evolution, according to which:
where the absolute velocity of the fluid at the rotor inlet
c4 is related to the isentropic enthalpy drop in the radial gap Δ
His(gap):
As in the previous cases, being initially unknown, the static pressure at the rotor inlet
P4, and hence the isentropic enthalpy drop in the radial gap, the absolute velocity
c4 could not be evaluated through Equation (30); an iterative calculation procedure was employed in this case, also adopting a first attempt value of 100 m/s for the absolute velocity
c4. The static temperature at the rotor inlet
T4 was then calculated as:
The static pressure
P4 could be instead obtained through the isentropic evolution from the nozzle exit (condition 3) to rotor inlet (condition 4):
The density of the fluid at the rotor inlet could be hence calculated by the application of the ideal gas law:
which, in turn, allowed calculation of the mass flow at the rotor inlet
G4:
where
A4 is the flow passage section at the rotor inlet, while the meridional component of the absolute velocity
c4m was obtained as:
The absolute velocity
c4 at rotor inlet was hence corrected on the basis of the mass-flow error (
G4 −
G5):
where
G5 is the mass flow through the outlet section from the rotor. The calculation procedure from Equation (29) to Equation (36) was repeated until the mass-flow convergence was obtained.
2.4. Analysis of the Rotor
The rotor is composed of a set of rotating blades that constitute a rotating ring of vanes, conceived to transform part of the fluid enthalpy and kinetic energy into mechanical work, which, in turn, is transferred to the rotating shaft. In a moving stage, multiple entropy generation mechanisms contribute to losses; the loss models mainly adopted in the scientific literature when managing the rotor of a radial turbine are: rotor incidence loss, passage loss, clearance loss, disc windage loss, and trailing edge loss [
18,
19,
20].
The rotor incidence loss is related to the incidence of the fluid flows approaching the rotor passage; according to the mostly employed model, described by NASA [
27], the enthalpy variation Δ
hin associated with the rotor incidence loss depends on the difference between the actual fluid-dynamic angle
β4,f and the optimum incidence angle
β4,opt:
where
w4 represents the relative velocity of the fluid at the rotor inlet; it is worth mentioning that the optimum angle
β4,opt is different from the geometric angle
β4,g due to the motion that the rotor induces in the flow approaching the blades. The optimum angle
β4,opt can be evaluated on the basis of the optimal tangential component of relative velocity
w4u,opt:
where
The optimal tangential component of the absolute velocity
c4u,opt in turn can be evaluated as a function of the peripheral linear velocity at the rotor inlet
u4 by means of an empirical equation proposed by Stanitz [
28]:
Hence, it results that the optimum angle
β4,opt can be evaluated as:
The rotor passage loss is a generic term aimed to quantify the losses due to friction and secondary flow processes that occur in the passage through the rotor vanes; the two contributions are usually incorporated into a single enthalpy variation Δ
hp since there is currently no way to isolate and measure their effects separately. In [
21], the rotor passage losses are quantified through:
where
Lhyd,R and
Dhyd,R are the hydraulic length and diameter of the rotor, respectively
, β5 and
b5 are the geometric angle and the blade height at rotor outlet, respectively, and
chrot is the rotor blade chord, which, according to [
21] can be approximated as:
where
Kp is a coefficient which, as suggested in [
21], should be set to 0.11 on the basis of some experimental data.
The clearance loss is the loss related to the fluid leakage from the rotor blade through the clearance gap between the rotor and its shroud. This loss seems to be affected to a greater extent by the radial clearance
εr than by the axial clearance
εx, and there appears to be a cross-coupling effect between the two parameters. The authors evaluated the enthalpy variation Δ
hcl due to the rotor clearance loss as reported in [
21]:
where
NR is the number of blades in the rotor,
Cx and
Cr are geometrical parameters, and the three coefficients
Kx, Kr, and
Kxr should be set to 0.4, 0.75, and −0.3, respectively, as indicated in [
21], in agreement with the previously described influences of
εr and
εx on rotor clearance loss.
The rotor disc windage loss is an external loss that expresses the power loss due to friction between the back face of the turbine disc and the fluid entrapped in the interspace between the rotor and its housing. According to [
29], the enthalpy variation connected to the rotor windage loss Δ
hw can be modelled as:
where
is the average fluid density between inlet and outlet section of the rotor,
G5 is the turbine mass flow, and the coefficient
Kf, as described in [
30], depends on the Reynolds number evaluated at the rotor inlet (Section 4 in
Figure 3) and on the turbine geometry:
where
εb is the clearance between the back face of the turbine disc and its housing, and
b4 represents the blade height at rotor inlet.
Lastly, the trailing-edge loss [
31] is related to the sudden expansion of the fluid when it passes the trailing edge of the rotor. The expansion is due to the rapid flow section increment caused by the end of the rotor blades. It is worth pointing out that this loss becomes numerically relevant if the fluid velocity is high, i.e., if the relative Mach number
M5,rel (evaluated on the basis of the relative velocity at the rotor exit
w5) approaches 1. The model here adopted by the authors calculates the enthalpy variation Δ
ht related to the trailing-edge loss as:
where
P5 and
T5 are the static pressure and temperature at the rotor exit (Section 5),
M5,rel is the previously mentioned relative Mach number in Section 5,
k and
Cp are the isentropic coefficient and the specific heat, respectively (both evaluated at the temperature
T5), and Δ
Prel is the pressure drop caused by the sudden expansion, which, according to the model adopted [
31], is assumed to be proportional to the relative kinetic energy at the rotor exit:
where
r5s and
r5h are the shroud and hub radii at rotor exit, respectively (see
Figure 3),
NR the number of blades in the rotor, and
t the blade thickness. In the calculation performed on the rotor, the outlet static pressure
P5 is known, being part of the boundary conditions adopted (as resumed in Part 2 [
12]); the isentropic enthalpy drop in the rotor can be hence calculated as:
Thus, the actual enthalpy drop in the rotor Δ
Hre(rot) can be obtained:
where Δ
hrot is the sum of all the losses in the rotor (=Δ
hin + Δ
hp + Δhcl + Δhw + Δht) discussed from Equation (37) to Equation (49). The static gas temperature
T5 at the rotor outlet is hence:
and the consequent fluid density
ρ5 obtained from the ideal gas law:
The actual enthalpy drop in the rotor also allows for calculation of the relative velocity at rotor outlet
w5 through the application of the first law of thermodynamics between Section 4 and Section 5, in the relative reference system of the rotor:
where
w4 is the relative velocity at the rotor inlet,
u4 and
u5 are the peripheral linear velocities at the rotor inlet and outlet, respectively. Once calculated, the value of
w5 is updated in Equations (48) and (49), thus allowing an iterative solution. The meridional component of the velocity
w5 is:
where
β5,g is the geometric blade angle at rotor outlet (see
Figure 5). The mass-flow rate can be hence calculated:
which, as already mentioned, is employed in Equation (36) for the iterative correction of the absolute velocity at the rotor inlet
c4. Since all the losses were expressed in terms of enthalpy variations, it is possible to evaluate the total-to-static isentropic efficiency of the stage as follows:
where the ideal enthalpy drop Δ
Hid is calculated considering an isentropic expansion from the inlet conditions (
PT1,
TT1) to the rotor exit static pressure
P5:
where
2.5. Mechanical Friction Losses
A radial inflow turbine needs at least a journal bearing for shaft rotation and an axial thrust bearing that transfers the axial load of the turbine rotor to the machine frame [
32]. The power dissipation in the journal bearing is primarily dependent on bearing geometry, oil viscosity, rotational speed, and oil film thickness. Generally, the frictional power loss in journal bearings can be estimated through Petroff’s equation [
33] as:
where
µ is the dynamic viscosity of the oil, ω is the rotational speed,
L is the bearing length,
D is the bearing diameter, and
h is the oil film thickness. The bearing diameter
D was obtained by designing the shaft for infinite life according to the procedure described in [
33], applying the Goodman criterion with a safety factor of 8. The material used for the shaft is 40 NiCrMo 4430 steel, tempered at 540 °C, characterized by yield strength 1080 MPa, and an ultimate tensile strength of 1170 MPa. The oil chosen for this application is Castrol 5W40 and its properties are evaluated at the working temperature of 90 °C. Finally, bearing length
L was determined referring to available data for the
L/D ratios applied in similar applications.
According to Petroff’s equation, the bearing diameter is the largest influencing factor on power dissipation, followed by the rotational speed. However, decreasing the shaft diameter is not advisable because of the rotor dynamic issues.
The frictional power loss thrust bearings can be described by a modified Petroff’s equation as follows:
where
rin,t denotes the inner radius, determined by shaft diameter
D;
rext,t is the outer radius calculated by making reference to the
rext,t/
rin,t ratios used in similar cases; and
εth is the axial clearance, for which a value of 0.095 mm was assigned according to [
32]. Petroff’s equation reveals that the power dissipation is highly dependent on geometry, as for the journal bearing. Larger bearing contact surfaces result in higher power losses.