Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Experimental Thermal and Fluid Science 30 (2006) 427440 www.elsevier.

com/locate/etfs

Experimental and numerical analyses to enhance the performance of a microturbine diuser


Ernesto Benini *, Andrea Toolo, Andrea Lazzaretto
Department of Mechanical Engineering, University of Padova, Via Venezia, 1, 35131 Padova, Italy Received 8 April 2005; accepted 16 September 2005

Abstract This paper describes design and o-design behavior of a centrifugal compressor of a 100 kW gas turbine used for small scale power generation and establishes the guidelines to improve diuser performance. The rst part of the paper deals with the experimental and numerical tests on the overall machine: An extensive series of tests at dierent operating points and rotational speeds is performed using steady probe measurements at impeller exit and diuser exit; the numerical model features a mixing plane at impellerdiuser interface and therefore neglects the eect of unsteadiness due to rotorstator interaction. In the second part of the paper, the true time-dependent rotorstator interaction is investigated by means of a numerical model where a sliding mesh technique is adopted instead. The unsteady results are then processed and compared with the computed steady ow in the diuser. Finally, the geometry of the compressor diuser is optimized using an evolutionary algorithm coupled with a CFD code. 2005 Elsevier Inc. All rights reserved.
Keywords: Optimization; Evolutionary algorithms; Microturbines; Centrifugal compressors; Diusers

1. Introduction Microturbine centrifugal compressors require very compact diusers which must operate at the highest eciency while achieving an adequate pressure recovery and ow turning before the air enters the combustion chamber. Most present design congurations make use of a twostage vaned diuser [1]: the rst radial row is followed by a 90 annular bend that conveys the ow to an axial deswirl cascade. Little information regarding design guidelines is provided in the open literature to help accomplish the design objectives, and the diuser apparatus is traditionally designed following very basic rules [2]. Two major issues have to be considered for an ecient design: the most important of the two deals with the radial cascade, which is in fact the most critical because of the strong diusion that occurs and because of the interaction

Corresponding author. Tel.: +39 049 8276767; fax: +39 049 8276785. E-mail address: ernesto.benini@unipd.it (E. Benini).

with the impeller. The other is related to the design of the annular bend and the deswirl cascade: within these components the ow must not generate excessive losses (especially those originating from wall boundary layer growth and secondary ow development) and must leave the blade row with low level of swirl (usually not greater than 1525). In particular, the design of the radial cascade is dicult and involves a lot of designer expertise. The ow leaving the impeller is fully three-dimensional, featuring highly non-uniformities between the hub and the shroud and in the circumferential direction. As a result, a very complex and time-dependent ow eld usually occurs in the region between the impeller tip and the diuser throat due to secondary ows that develop within the machine. This aspect has been documented by several researchers (see, among others [3,4]). Owing to this complexity, some authors have underlined the weak points of a simple approach that attempts to describe the ow entering the diuser without the direct eect of impeller interaction [57]. On the other hand, with the help of both experimental and numerical simulations, other researchers have indicated that the

0894-1777/$ - see front matter 2005 Elsevier Inc. All rights reserved. doi:10.1016/j.expthermusci.2005.09.003

428

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

Nomenclature c chord length, m cn,i ith coecient of Bezier polygon Cp = p3 p2/p02 p2 pressure recovery coecient of diuser F vector of objective functions (f1, f2) g gap, m M Mach number _ m mass ow rate, kg/s n rotational speed, rpm p pressure, Pa R radius, m r, rh, z cylindrical coordinates, m t, s time, s T time period, s T temperature, K x vector of decision variables x coordinate along chord, m gis total-to-total isentropic eciency W generic ow quantity e0 W unsteady component of the generic ow quantity W h average value of the generic ow quantity angle between prole leading and trailing edges, measured in the tangential direction x = p03 p02/p02 p2 aerodynamic loss coecient of diuser Subscripts 0 total 1 compressor inlet 2 impeller outlet, diuser inlet 3 diuser outlet ax axial diuser clocking circumferential clocking hub hub le leading edge ps pressure side rad radial diuser ss suction side te trailing edge
0

circumferential non-uniformities are less important than those occurring from hub to shroud in the neighborhood of the best eciency working point [3,8]. Moreover, many works report that the uctuations of the thermo-uid quantities as well as of performance parameters decay very rapidly in the diuser [912]: the insensitivity of diuser performance to the incoming pulsating ow justies the highly successful diuser designs that do not account for rotorstator interaction. This fact has recently encouraged some researchers to study the ow inside diusers (without the impeller eect) using computational uid dynamics (CFD) [13,14], and to develop methodologies to optimize diuser performance. Regarding the latter, Benini and Tourlidakis [15] used a Pareto genetic algorithm and CFD to optimize the shape of a channel diuser. Zangeneh et al. [16] used a 3D inverse design technique to improve the pressure recovery of a vaned diuser for a given impeller. This paper deals with performance evaluation and optimization of a centrifugal compressor diuser used in a small gas turbine (100 kW). The paper is virtually divided into three parts. The rst refers to the experimental and numerical investigation on the overall compressor and diffuser in both design and o-design conditions. The experimental tests are performed according to both ASME and UNI-ISO standards. In the second part, CFD is exploited to simulate, visualize and analyze the complex ow generated by the rotorstator interaction, with particular emphasis on the unsteady behavior of the vaned diuser. The third part deals with the numerical constrained optimi-

zation of the diuser apparatus (i.e. the radial and deswirl cascades) for maximum aerodynamic eciency and pressure recovery. The investigated compressor is part of the microturbine SOLAR T62, which is widely used as an auxiliary power unit (APU) in military helicopters and as a ground power unit (GPU) in small light helicopters. The turbine studied in this work is actually the Titan T62T32, a version conceived for continuous operation that can also be used for ground electric power generation. It consists of a onestage centrifugal compressor mounted back-to-back with a one-stage radial inow turbine wheel and an annular reverse-ow combustion chamber (Fig. 1). At the design point, the microturbine develops approximately 100 kW shaft power at the rotational speed of 60,000 rpm. Under these conditions, the pressure ratio is 3.5, the air mass ow rate is 1 kg/s, turbine inlet temperature is 788 C and the exhaust gas temperature is 560 C. 2. Investigation on overall compressor and diuser performance 2.1. Experimental apparatus In order to measure the overall compressor and diuser performance, a test rig was set up where the compressor was driven by the microturbine, as in normal engine operation. A sketch of the test rig is given in Fig. 2. The rig consisted of a test-bed where the microturbine was mounted and connected to an eddy-current brake, which measured

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

429

Fig. 1. Cutaway drawing of the SOLAR T62T32 microturbine.

the torque developed by the engine and the rotational speed of the output shaft. Since the compressor intake is not straight, an inlet pipe (Fig. 3) was built to convey the mass ow rate entering the compressor. The mass ow rate was estimated by measuring the air velocity in the pipe (using a Pitot probe) and air density, as suggested in [17,18]. The mass ow rate was changed by means of a gate valve positioned at the beginning of the inlet pipe. A set of calibrated probes for pressure and temperature measurements were placed upstream and downstream of the compression stage, as well as in between the impeller and diuser. In all, the following measurement probes were used (see Fig. 2): 1. static pressure probes for p1; 2. rack of total pressure probes for p01; 3. total temperature probes for T01;

4. 5. 6. 7. 8. 9. 10.

total temperature probes for T02; rack of total pressure probes for p02; static pressure probes for p2; rack of total pressure probes for p03; total temperature probes for T03; static pressure probes for p3; pitot probe for mass ow rate.

Details on characteristic dimensions and shape of pressure and temperature probes at impeller and diuser exits can be found in Table 1 and in Figs. 4 and 5. Note that, according to [19], the internal cone angle of total pressure probes results in a sensitivity angle of about 25 when ow is not aligned with probe axis. This made it possible to determine compressor performance over the entire operating range. The resolution of pressure and temperature signals is 100 Pa and 0.05 K, respectively. Calibration of

430

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

Fig. 2. Sketch of compressor test rig.

Fig. 3. View of complete compressor test rig. Table 1 Probe characteristics Impeller exit Static pressure Total pressure No. of holes Diameter No. of probes External diameter Internal diameter Internal cone angle No. of probes External diameter Internal diameter Type Insulation 4 1 mm 2 1.2 mm 0.75 mm 30 2 2 mm 1.5 mm K PTFE Diuser exit 4 1 mm 4 racks of 2 1.2 mm 0.75 mm 30 4 2 mm 1.5 mm K PTFE Fig. 4. Pressure and temperature probes at impeller exit.

Total temperature

pressure transducers and thermocouples was performed using instruments having superior metrological characteris-

tics. All the probes showed a highly linear behavior within the required range of measurement. The uncertainty on the mass ow measurement was determined with the help of the standard UNI EN ISO 5167 [20]. The uncertainties of the various terms are: uncertainty at the 95% condence level in input power measurements is about 0.3%; uncertainty of the electromechanical eciency is about 0.9% when input power is 100 kW and electromechanical eciency is 75%; uncertainty of the mass ow rate is about 0.9%; uncertainty of the pressure ratio is about 0.6%; uncertainty of the isentropic eciency is

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

431

nomena, that is when periodic noise and intense vibrations occurred. Diuser performance parameters were also determined at the rotational speeds dened above. 2.2. CFD simulations In the numerical study of the compressor, a steady state analysis was performed using a mixing plane approach with a single rotating reference frame. This implies that only one impeller channel and one third part of the diuser were modeled for simulation, since periodic boundaries were adopted. Using such an approach, the governing equations are solved in a reference frame that rotates at the rotational speed of the impeller. The interface between impeller and diuser was modeled using a mixing plane: thermo- and uid-dynamic quantities are averaged in the pitchwise direction through the mixing plane, whereas their actual distribution is maintained in the axial direction. The outlet boundary was located at diuser exit. Structured single-block H-type grids were used to mesh both rotating and stationary blade passages (Fig. 6). The overall grid consisted of 160,342 nodes which were partitioned in the following way: Impeller 28,353 nodes (18%), diuser 131,989 nodes (82%). For simplicity, the impeller was modeled without tip clearance. A limited number of grid sensitivity studies were carried out to ensure a satisfactory accuracy of the ow solver. For this purpose, the compressor performance map was calculated with the baseline grid described above, as well as with two other grids: the rst was coarser (approximately 100,000 nodes) and the latter was ner (approximately 250,000 nodes). Although the results are not reported here for brevity, the sensitivity analysis showed that the baseline grid featured a better capability, with respect to the coarser one, in capturing

Fig. 5. Pressure and temperature probes at diuser exit.

about 1.2%; uncertainty of the compressor total eciency is about 0.9%. All measuring data were collected by a data logger into a computer that evaluated the performance parameters. As a reference for the determination of pressure ratio and isentropic eciency (total-to-total), the ambient pressure and ambient temperature were measured in the test room. Ambient humidity was registered as well. Pressure ratio and isentropic eciency were evaluated from surge to choke at dierent rotational speeds (100%n, 90%n, 80%n, 70%n). Surge was determined by identifying the surge phe-

Fig. 6. Grid used in CFD calculations.

432

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

the compressor characteristic curve at part load. On the other hand, no noticeable improvement with respect to the baseline conguration was found using the ner grid. This agrees with the statements reported in [21] and with the results published in [22]: In particular, in the latter paper the authors demonstrated that most of the important eects in centrifugal compressors (i.e. those related to overall performance) may be captured using a coarse grid of only 30,000 nodes, and that nearly all of the results regarding eciency and pressure rise agree well with respect to measured values. Therefore, all the results presented here were obtained using the baseline grid. Three-dimensional steady-state Reynolds-averaged NavierStokes equations were solved using Fluent5.4 code by Fluent, Inc. The uid was supposed to be a compressible ideal gas with constant specic heat capacities. A standard ke model [23] was used to account for turbulence in the mean ow and a wall function approach was chosen to solve the boundary layer. All walls (moving and stationary) were treated as hydraulically smooth and adiabatic. Two sets of boundary conditions, corresponding to the nearchoke and near-stall conditions, were considered. When operating conditions close to choking were to be analyzed, the measured values of the total pressure and total temperature were applied at impeller inlet; at diuser exit, the measured static pressure was instead prescribed: As a result, the overall mass ow rate of the compressor was estimated. When the working point approached surge, the values of the mass ow rate and total temperature were applied at the inlet, while the measured static pressure was applied at diuser exit. All the calculated quantities were based on their mass-average values. Using such a steady state approach it was possible to simulate working conditions even beyond the compressor stall point obtained experimentally (i.e. at reduced mass ow rates). The last computation for which the CFD code was able to converge was considered a numerical estimation of the stall point because the unsteady calculations performed beyond that point featured perceivable but unstable separations. 2.3. Experimental and numerical results The characteristic curves of the overall compressor, obtained both experimentally and numerically, are shown in Fig. 7. The pressure ratio p03/p01 and isentropic eciency gis are plotted as functions of the corrected mass ow rate at four values of the rotational speed: n, 0.9n, 0.8n and 0.7n. The experimental test showed that the compressor has quite a narrow operating range at all rotational speeds, and that the normal working point is very close to the choke line, the corrected mass ow being 1.033 kg/s and the pressure ratio 3.6. In this condition, the isentropic eciency is about 0.7. This operating point presumably gives adequate margins against the occurrence of compressor surge without heavy drawbacks on the eciency. At reduced rotational speeds, the pressure ratio curves atten out and suggest how careful the operation in these condi-

Fig. 7. Comparison between experimental and computed compressor characteristics.

tions should be in order to avoid the occurrence of ow instabilities. In some cases, signicant dierences were registered in the code computing accuracy concerning pressure ratio and eciency. At nominal rotational speed, predictions of the pressure ratio are excellent over the whole operating range, while those regarding the eciency are less accurate, the maximum discrepancy being in the order of 4%. At reduced rotational speeds, computed values of the eciency are quantitatively better while those of pressure ratio are worse (i.e. the code underestimates the pressure rise). In fact, as the rotational speed reduces, the computed characteristic curves are shifted toward lower values of the mass ow rate, i.e. the code found some diculties in capturing the choke condition. While examining these results, however, the uncertainty on experimental data as well as the limitations of the steady state approach and of the turbulence model, especially at part load, must be properly taken into account: the use of the steady-state approach, in particular, is known to give misleading results when the operating conditions are very far from the nominal one and strong recirculations at impellerdiuser interface are usually observed. In this case, however, the use of a mixing plane still gives acceptable results because of the narrow operating range of the compressor. Also, the authors believe that some of the discrepancies could be explained with the absence of the tip clearance in the numerical simulations, which would limit the maximum mass ow rate to some extents and would contribute to a reduction in the total pressure ratio and eciency. The characteristic curves of the diuser are reported in Fig. 8. They show both the experimental and computed values of the pressure recovery coecient Cp and the aerodynamic loss coecient x as functions of the corrected mass ow rate. At the nominal speed, the measured pressure recovery coecient decreases from 0.5 to 0.45 as the mass ow rate increases from stall to choke (i.e. as the angle of the absolute velocity with respect the tangential direction

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

433

Fig. 8. Comparison between experimental and computed diuser characteristics.

increases from 19.4 to 19.9). In such conditions, the aerodynamic loss coecient increases from 0.45 to 0.51. The same tendency is qualitatively observed at reduced speeds, even though the pressure recovery coecient is higher at part load, because the dynamic pressure at impeller exit decreases much more than the diuser pressure recovery. On the other hand, the aerodynamic loss coecient reduces slightly at lower mass ow rates due to the fast decrement of the total pressure loss within the diuser. Because of the lack of measurement probes in the annular bend, it was not possible to isolate the eect of the radial and axial blade rows on overall diuser performance and, therefore, to establish whether or not the axial deswirl cascade has a negative eect on compressor eciency and stability. The agreement between the numerical and experimental results is good at design nominal speed, whereas areas of relatively poor code accuracy were found at reduced rotational speed. Again, this can be justied if the simplications of the numerical model at impellerdiffuser interface are taken into account. In any case, the overall results suggest that the CFD model is suciently accurate to give realistic indications on diuser performance within the overall compressor operating range. 3. Numerical analysis of impellerdiuser interaction 3.1. Objectives and approach Viscous and potential eects of rotorstator interaction are comparable in high-speed centrifugal compressors, since the mixing process that rotor blade wakes undergo is very fast, and the radial gap between rotors and vaned

diusers is usually very small. As a consequence, a mutual interaction occurs between the components. The inuence of the impeller on diuser ow is mainly characterized by viscous eects caused by rotor wakes, while the inuence of the diuser on the impeller ow is mainly caused by potential eects [1,24,25]. However, a number of experimental works [3,8,12,26] have shown that the circumferential ow non-uniformity at impeller exit mixes out very rapidly near the design point, so diuser ow can be very well approximated as steady. Dawes [5] and Yamane et al. [27] compared a steady approach featuring a mixing plane between impeller and vaned diuser and the corresponding fully-coupled unsteady approach. They clearly showed the inuence of unsteady eects in impeller and diffuser, in particular the unsteady eects due to the highly distorted impeller ow eld (both circumferentially and axially) and those due to the wakes released by impeller blades. In this work, two approaches were investigated: the unsteady-fully-coupled and a steady-decoupled approach, in which each blade row is treated separately by steady computations and ow quantities at the interface are averaged in time and in the circumferential direction (while preserving their spanwise distribution). The latter approach gives satisfactory results provided that a proper averaging is carried out, and this may be not the case when the spacing between the blade rows is small. An alternative strategy is the frozen rotor approach (see, among others [28]), in which steady calculations are performed in a number of xed impellerdiuser positions. However it was not explored here because its features are taken from both the other approaches and do not help to achieve either accuracy or computational speediness.

434

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

The objective of the work is twofold: (i) to achieve a better understanding of the unsteady ow phenomena involved during the interaction; (ii) to assess a quantitative measure of ow unsteadiness within the diuser, in order to verify how much diuser performance is aected by the presence of the impeller. The latter objective is fundamental in the optimization of the diuser apparatus, described in the next section, where several time consuming CFD calculations are needed to achieve the nal solution. A fully realistic diuser can be modeled without the upstream impeller provided that the boundary condition at the interface accurately simulates the presence of the impeller. A typical diculty associated with multi-row simulations is that each blade row generally has a dierent number of blades, and that the ratio between the rotor and stator blades (or its inverse) is not an integer value. In our case, due to the lack of periodicity in the blade number of compressor components, the number of blades of the diuser was modied (12-blades radial and 36-blades deswirl), in order to simplify the ow domain and reduce the computational eort. As a result, it should not be expected that the computed solution represents accurately the real ow within the original conguration. However, the purpose here was much more focused on assessing a methodology for studying the interaction rather than simulating the real ow. The resulting ow domain was divided into three blocks, each representing a quarter of the real physical domain, for the impeller, the radial and deswirl diusers, respectively. The assembled grid consisted of about 100,000 nodes. The impeller was modeled without tip clearance. The unsteady statorrotor simulations were carried out using the CFD code FluentTM, by Fluent Inc., where a sliding mesh technique was utilized. The unsteady 3D Reynolds-averaged NavierStokes (RANS) equations for a compressible ideal gas were solved along with a Spalart Allmaras turbulence model [29]. Standard wall functions were used to link the solution variables at the near-wall cells and the corresponding quantities on the wall. Boundary conditions were imposed as follows: the total pressure and total temperature were applied at impeller inlet (p01 = 101325 Pa, T01 = 288.1 K), where the ow was supposed to be swirl-free; a constant value of the static pressure was maintained at the outlet of the deswirl diuser (p3 = 193913 Pa). 3.2. Results The unsteady computations were carried out using a time step Dt = 4 106 s, which corresponds to 1.47 of impeller rotation. The chosen time step was the maximum that made it possible to capture the ow unsteadiness with reasonable accuracy avoiding the increase of the computational eort beyond unacceptable limits. An unsteady run required about 50 h to reach a periodic solution on a Workstation AlphaServer ES40, clock 667 MHz, 1.5 GB RAM. The periodicity of the pressure signal at a point

located between impeller and diuser blades was considered as a convergence criterion, which was typically satised after one to two complete impeller revolutions. The time average of the unsteady solution was calculated by means of an in-house post-processing tool in order to isolate the unsteady components of the ow quantities. The unsteady component of the generic quantity W (r, h, z, t) was calculated as follows: Wr; h; z; t Wr; h; z e0 W r; h; z; t Wr; h; z where 1 Wr; h; z T Z
0 T

Wr; h; z; tdt

The instantaneous contours of pressure and Mach nume0 ber and of their unsteady components ~0 and M are p reported in Figs. 9 and 10, respectively, on a plane which cuts blade passages at impeller exit midspan. These plots correspond to various instants in time during the period T, i.e. the time required for the rotor blade to cover the distance corresponding to one rotor pitch. Note that quantities are time-averaged with reference to the absolute frame in the diuser, whereas, quantities are time-averaged with reference to the rotating frame in the sliding mesh portion including the impeller. As a consequence, a discontinuity at the impellerdiuser interface appears in Figs. 9(b) and 10(b). The jet ow leaving the impeller is periodically cut by the diuser leading edge, and this causes periodic pressure uctuations on both impeller trailing edges and diuser leading edges. Thus, the largest part of ow unsteadiness comes from the potential eect, which follows the periodic cycles of rotating blade positions relative to the stationary one, and is conned in the semi-vaneless gap between the impeller and radial diuser. The magnitude of such unsteadiness is in the order of 10% for both the pressure and Mach number, the core being located very close to the middle of the gap between the rotating and stationary components. As the ow mixes out in the diuser, the unsteadiness reduces rapidly and the ow becomes nearly steady. It is worth noting that the ow within the radial diuser, in particular toward the trailing edge, is almost insensitive to the unsteadiness generated by the impeller. However, quite large uctuations of the Mach number can be identied in the diuser along the surfaces of the radial cascade, both on the pressure and suction sides. These uctuations are probably due to the fact that the stagger angle of the radial cascade is not properly matched with the angle of the ow leaving the impeller. Therefore, the ow accelerates or decelerates according to the reciprocal position between the impeller and diuser blades, i.e. according to both the interception of the jets from the impeller by the radial prole itself (wake eect) and the oscillations of the pressure eld in the gap between impeller and diuser cascades (potential eect).

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

435

Fig. 9. Instantaneous contours of pressure at diuser midspan for dierent impeller blade positions (a) and instantaneous contours of unsteady pressure (b).

Fig. 10. Instantaneous contours of Mach number at diuser midspan for dierent impeller blade positions (a) and instantaneous contours of unsteady Mach number (b).

The unsteady performance coecients (x and Cp) of the overall diuser apparatus were calculated as well, and their values are reported in Fig. 11 as a function of the impeller revolutions counter. The mass-ow weighted average values were calculated and made dimensionless with respect

to their time average value: it can be noted that the uctuations of x are less than 0.5% and those of Cp are even less. The time averaged values of x and Cp were then compared with those obtained from a steady simulation regarding the diuser alone, i.e. without the impeller. This

436

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

where x is the vector of design optimization parameters, that is the decision variables of the problem. The chosen objectives cannot be satised simultaneously by a single design, since maximum pressure rise is achieved through high aerodynamic loading on blade proles, resulting in higher total pressure losses, whereas minimum total pressure loss is obtained using low solidity cascades to minimize friction losses, without altering too much ow tangential direction. Thus, Pareto optimality is used to rank the solutions examined during the optimization process and to obtain the true trade-o solutions between the two objectives (Pareto front). A special evaluation method is applied in order to improve the search capabilities of the MOEA and to spread the optimal solutions as uniformly as possible along the Pareto front [30]. 4.2. Denition of design parameters
Fig. 11. Comparison between steady and unsteady performance coecients of the diuser apparatus.

simulation was carried out using the same solver, settings and grid: the values of the relevant quantities of the unsteady solution, averaged with respect to time and mass ow rate, were assigned as boundary condition at diuser inlet. In this way the dynamic eect of the impeller was neglected. In particular, time- and circumferentially-massaveraged values of the total pressure and total temperature were applied at diuser inlet; time- and mass-averaged value for the static pressure was instead xed at diuser outlet. The results of the steady calculation are given in Fig. 11 and compared with those of the unsteady computation. The dierences are virtually negligible: 1.9% in x, 0.6% in Cp. These results apparently demonstrate that, from the point of view of diuser performance, the presence of the impeller can be reproduced by assigning averaged and steady boundary conditions at impeller outlet. 4. Optimization of diuser performance 4.1. Objectives and approach In this section, a numerical multi-objective optimization of the radial and deswirl cascades of the centrifugal compressor is accomplished through an iterative procedure based on the combination between a Multi-Objective Evolutionary Algorithm (MOEA) and a CFD model of the diffuser. The aim is to develop a set of diuser designs achieving maximum pressure rise (maximum Cp) and minimum total pressure loss (maximum 1 x) at the design condition (see previous section). These designs have also to t into the radial and axial sizes of the original one, perhaps adjusting the radius between the radial and the deswirl sections. The optimization problem is to maximize the two-objective function: F x f1 ; f2 C p ; 1 x 3

Since the optimal designs have to t into the overall size of the original one in the meridional plane, the following dimensions are chosen as optimization parameters (see Fig. 12): the radius Rhub of the arc linking the radial and the deswirl sections of the diuser in a meridional plane; the radial coordinate Rle,rad of the radial prole leading edge; the radial clearance gte,rad between the radial prole trailing edge and the radius linking the two sections; the angle hrad between the radial prole leading and trailing edges, with respect to the tangential direction; the axial clearance gle,ax between the deswirl prole leading edge and the radius linking the two sections; the angle hax between the deswirl prole leading and trailing edges, with respect to the tangential direction; the tangential clocking hclocking between the leading edges of the radial and the deswirl proles. The number of blades of both radial and deswirl sections (12 and 36, respectively) is the one used in the previous section. The shape of blade proles is described using two Bezier parametric curves (one for the pressure side and one for the suction side). The non-dimensional Cartesian coordinates of each curve x, yps and x, yss are dened by n + 1 control points constituting the Bezier polygon according to the following expression: n X cn;i ti 1 tni fxi ; y ps;i ; y ss;i g 4 fxt; y ps t; y ss tg
i0

where t 2 [0, 1] is the non-dimensional parameter and cn,i = n!/(i!(n i)!). The n + 1 control points coordinates xi, yps,i, yss,i are dened as follows: xi 0; 0; x2 ; . . . ; xn2 ; 1; 1 y ps;i 0; dle ; . . . ; y cl;i di ; . . . ; dte ; 0 y ss;i 0; dle ; . . . ; y cl;i di ; . . . ; dte ; 0 where dle and dte x the thickness of the leading and trailing edges, respectively, and ycl,i and di are the actual optimization parameters for blade shape geometry. This parameterization is inspired by the well-known practice of 5

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440 Table 2 Ranges of optimization parameters Variable Rhub Rle,rad gte,rad hrad y1,rad y2,rad y3,rad d1,rad d2,rad d3,rad gle,ax hax y1,ax y2,ax d1,ax d2,ax hclocking Unit mm mm mm mm Original 9.5 84 3.5 40 0.02 0.055 0.06 0.03 0.065 0.055 2.5 5 0.12 0.12 0.055 0.055 6.1 Min 8 82 2 25 0.015 0.05 0.04 0.025 0.055 0.035 1 2 0 0 0.02 0.02 0 Max 12 86 6 35 0.025 0.06 0.08 0.035 0.075 0.075 3 8 0.2 0.2 0.1 0.1 10

437

4.3. Results and discussion The optimization algorithm was run for 40 generations with a population size of 50 individuals. The most important design parameters and the corresponding objective function values for the last generation of solution are presented in Fig. 13. Note that performance indexes of the original design are too low (Cp = 0.45 and 1 x = 0.5) to appear in the Figure. It is apparent that all the individuals fall in a narrow strip of the plane dened by the two objective functions Cp and 1 x. As a matter of fact, the conict between the objectives seems of little account, but this is simply because of the tight ranges of variation assigned to the optimization variables. The Pareto front is made of only two individuals, one of them maximizing pressure recovery (marked in red) and the other minimizing total pressure losses (marked in blue). These optimal solutions are compared to the original design in Fig. 13.
Fig. 12. Denition of optimization parameters.

superimposing a thickness distribution to a chamber line and avoids the generation of intersecting pressure and suction side curves. The geometry of the diuser radial and deswirl blades is described here using seven and six control points, respectively (6 and 4 decision variables). The total number of optimization parameters is therefore 17. According to the geometrical constraints, the ranges of variation chosen for most optimization variables are narrow and centered around the corresponding values of the original design. The main exception to this criterion is represented by the value of hrad, which is varied in a range that does not include the original value. The shape of the deswirl prole is also allowed to vary more freely than that of the radial prole. The ranges of variation for all the optimization variables are summarized in Table 2.

4.3.1. The radial prole The shape of the optimal radial proles is very similar to the original one because of the limits imposed on the variation of its Bezier control points. On the other hand, the stagger angle, which is varied in a wider range, seems to be the most signicant design parameter. The optimized solutions have a much lower value of hrad than the original design (40). This results in higher pressure recovery, because of reduced tangential velocity components, as well as in lower total pressure losses because of better incidence angles and less friction on shorter proles. hrad being approximately the same, the actual conict between the two objectives owing to two opposite trends toward shorter proles to minimize losses and toward longer proles to maximize pressure rise. The radial coordinate of the leading edge Rle,rad tends toward its maximum value to shorten the prole and to allow an initial pressure recovery without blade friction, that is with lower total pressure losses. This fact agrees with

438

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

Fig. 13. Results of optimization.

the experimental results published in the literature about high-speed compressors (see, among others [31]). In fact, as the absolute Mach number of the ow leaving the impeller is quite high, a longer vaneless gap is needed to reduce the Mach number level, and therefore losses, before entering the vanes. The optimized proles are also less thick than the original ones, resulting in a higher pressure rise. Even though ow separation near the trailing edge is more likely for thinner proles, it does not happen in this case due to the small ranges of variation imposed on the shape of the prole. 4.3.2. The meridional channel The radius Rhub linking the radial and the axial section of the meridional channel tends to its maximum value for both the optimal designs. This is reasonable, since the loss related to secondary ows in the bend and downstream of it is reduced. The enlargement of the gap between the radial coordinate of the trailing edge and the beginning of the curvature may also be responsible for this reduction. 4.3.3. The deswirl prole The orientation of the deswirl prole in the optimized designs cannot be analyzed separately from the ow deviation imposed by the radial cascade of the diuser. Since hrad is lower than in the original design, the tangential velocity of the ow entering the deswirl cascade is lower because of the conservation of the tangential momentum. On the other hand, the meridional velocity is also lower because of the conservation of meridional momentum, since Rhub is larger in the optimized designs. The latter eect prevails over the former, and the result is a stagger

angle, measured by hax, that is higher than in the original design. Even though the vanishing of the tangential component through the deswirl cascade would result in the maximum pressure recovery, the exit angle of the optimized designs is far from the value that would fulll this condition. Perhaps the chord of the prole is too short to accomplish the ideal deviation at the expense of a negligible increase in total pressure losses. Finally, the optimized clocking hclocking is achieved when the wake of the radial prole wraps one of the deswirl proles. 5. Practical signicance/usefulness The methodology described in the rst part of the paper signicantly reduces the computational eort required to perform CFD analyses on rotating machinery featuring high-speed ows, compressibility issues, rotor/stator interaction and problems related to the denition of proper boundary conditions. This can be achieved without compromises on the accuracy of the numerical results, as certied by the validation presented. The ultimate goal is to use such approaches to tackle complex optimization problems using advanced mathematical techniques. From this point of view, evolutionary algorithms show very attractive potentials in the exploration of wide search spaces with many decision variables and complex and conicting objective functions. 6. Conclusions In this paper, experimental and numerical analyses were used to achieve performance enhancements of a microturbine diuser.

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440

439

First of all, a methodology for testing the centrifugal compressor was presented. A test rig was set up and equipped with pressure and temperature probes at impeller inlet and outlet and at diuser exit. The tests were carried out according to both ASME and UNI-ISO standards. A numerical model based on 3D CFD was also carried out and validated against the experimental data. The numerical results regarding the pressure ratio at the nominal speed agree with the experimental data; those concerning the isentropic eciency show poorer agreement (computed eciency is higher). At reduced speeds, the numerical model overestimates the pressure ratio to some extent, whereas calculated eciency is much closer to the measured one; at the same time, the mass ow rate at choking is a little lower than the one observed experimentally. These facts may be explained by the absence of the tip clearance in the numerical model of the impeller (see also the discussion in Section 2.3). The diuser maps showed that the pressure recovery and aerodynamic losses are apparent functions of the mass ow rate for a given compressor rotational speed. The absence of measurement probes in the space between the radial and deswirl blade rows did not make it possible to describe the behavior of each cascade, even though it is known that the actual pressure recovery in the last row is not likely to be high; its function is mainly to remove swirl before the ow enters the combustor. However, the eect of the axial deswirl diuser should be investigated further in order to establish its inuence on compressor eciency and stability. The numerical results showed that the steady approach is suciently accurate to predict the characteristics of the diuser, at least for the nominal rotational speed. At reduced speeds, in particular at part loads and near compressor stall, somewhat poor agreement with the experimental data suggests that diuser performance could be signicantly inuenced by the jet-wake and recirculation ow structures which the CFD model, being based on a mixing plane approach, obviously was not able to capture. In the second part of the paper, two approaches for the analysis of impellerdiuser interaction in the centrifugal compressor stage were examined. The rst approach was based on the fully-coupled unsteady solution of the ow eld; the latter assumed time- and space-averaged boundary conditions at the interface between the impeller and diffuser with which a steady and decoupled solution was obtained. The unsteady simulation made it possible to analyze and understand the details of the main sources of ow eld uctuations. The amplitudes of these uctuations are remarkable only in the semi-vaneless gap, whereas the diffuser blade channel is not substantially involved in these phenomena. This fact is conrmed by the agreement with the results of the steady simulation, performance indexes being virtually identical. The results of both approaches highlight that some of the geometrical characteristics of the diuser are not properly matched to the ow leaving the impeller, leading to a poor overall diuser performance. In the third part of the paper, the diuser design was optimized focusing the attention on the cascade parame-

ters, and leaving the size of the meridional channel unchanged. The aim was to obtain the maximum pressure rise at the minimum total pressure loss. The conict between the two objective is minimal, owing to the tight constraints imposed to the chosen design variables. The most signicant dierences from the original design are the lower stagger angle of the radial prole, leading to higher pressure rises, and the lower chamber of the deswirl prole, which actually hardly deects the ow resulting in lower total pressure losses. The optimization of diuser performance focusing on the design point only is meaningful, since the operating range of this microturbine centrifugal compressor is very narrow. In o-design conditions, in fact, choking or stall occur before performance drops due to unsatisfactory design solutions. References
[1] C. Rodgers. Microturbine cycle options, ASME Paper 2001-GT-0552, 2001. [2] D. Japikse, Centrifugal Compressor Design and Performance, Concepts ETI Inc., Wilder, Vermont, 1996. [3] M. Inoue, N.A. Cumpsty, Experimental study of centrifugal compressor discharge ow in vaneless and vaned diusers, ASME Journal of Engineering for Gas Turbines and Power 106 (3) (1984) 455467. [4] D. Bonaiuti, A. Arnone, C. Hah, H. Hayami. Development of secondary ow eld in a low solidity diuser in a transonic centrifugal compressor stage, ASME Paper GT-2002-30371, 2002. [5] W. N. Dawes. A simulation of the unsteady interaction of a centrifugal impeller with its vaned diuser: ow analysis, ASME Paper 94-GT-105, 1994. [6] K. Sato, L. He. A numerical study on performance of centrifugal compressor stages with dierent radial gaps, ASME Paper 2000-GT462, 2000. [7] Y.K.P. Shum, C.S. Tan, N.A. Cumpsty. Impellerdiuser interaction in centrifugal compressor, ASME Paper 2000-GT-0428, 2000. [8] Y. Senoo. Vaneless diusers and vaned diusers. Flow in Centrifugal Compressors, VKI Lecture Series 1984-07, 1984. [9] M.V. Casey, M. Eisele, Z. Zhang, J. Gulich, A. Schachenmann. Flow analysis in a pump diuser, Part 1: LDA and PTV measurements of the unsteady ow, ASME Paper, FED Summer Meeting, Symposium on Laser Anemometry, 1995. [10] M.V. Casey, M. Eisele, F.A. Muggli, J. Gulich, A. Schachenmann. Flow analysis in a pump diuser, Part 2: validation of a CFD code for steady ow, ASME Paper, FED Summer Meeting, Symposium on Laser Anemometry, 1995. [11] F.A. Muggli, D. Wiss, K. Eisele, Z. Zhang, M.A. Casey, P. Galpin. Unsteady ow in the vaned diuser of a medium specic speed pump, ASME Paper 96-GT-157, 1996. [12] S. Deniz, E.M. Greitzer, N.A. Cumpsty, Eects of inlet ow eld conditions on the performance of centrifugal compressor diusers: Part 2Straight-channel diuser, ASME Journal of Turbomachinery 122 (1) (2000) 1121. [13] P. Dalbert, G. Gyarmathy, A. Sebestyen. Flow phenomena in vaned diuser of a centrifugal compressor, ASME Paper 93-GT-53, 1993. [14] E. Casartelli, A.P. Saxer, G. Gyarmathy. Performance analysis in a subsonic radial diuser, ASME Paper 98-GT-153, 1998. [15] E. Benini, A. Tourlidakis. Design optimization of vaned diusers for centrifugal compressors using genetic algorithms. AIAA Paper 20012583, 2001. [16] M. Zangeneh, D. Vogt, C. Roduner. Improving a vaned diuser for a given centrifugal impeller by 3D inverse design, ASME Paper GT2002-30621, 2002. [17] ASME PTC 10-1997. Performance test code on compressor and exhausters, ASME Code, 1997.

440

E. Benini et al. / Experimental Thermal and Fluid Science 30 (2006) 427440 [26] V.G. Filipenco, S. Deniz, J.M. Johnston, E.M. Greitzer, N.A. Cumpsty, Eects of inlet ow eld conditions on the performance of centrifugal compressor diusers: Part 1Discrete-passage diuser, ASME Journal of Turbomachinery 122 (1) (2000) 110. [27] T. Yamane, H. Fujita, T. Nagashima. Transonic discharge ows around diuser vanes from a centrifugal impeller, ISABE Paper 937053, 1993. [28] Z. Liu and D.L. Hill. Issues surrounding multiple frames of reference models for turbo compressor applications, in: Proceedings, 15th International Compressor Engineering Conference at Purdue University, West Lafayette, 2000. [29] P.R. Spalart, S.R. Allmaras. A one-equation model for aerodynamic ows, AIAA Paper 92-0439, 1992. [30] A. Toolo, E. Benini, Genetic diversity as an objective in multiobjective evolutionary algorithms, Evolutionary Computation 11 (2) (2003). [31] R.H. Aungier, Centrifugal Compressors a Strategy for Aerodynamic Design and Analysis, ASME Press, New York, 2000, p. 2.

[18] UNI 10727. Fluids owrate in closed circular conduitsvelocity measurement method on one point of section, Italian Standard, 1998. [19] D. Japikse. Advanced experimental techniques in turbomachinery, Principal Lecture Series No. 1, Concepts ETI, Inc. Norwich, 2000. [20] UNI EN ISO 5167, Measurement of uid ow by means of pressure dierential devices, European Standard, 1997. [21] J.D. Denton. Turbomachinery Design Using CFD, AGARD Lecture Series 195, 1994. [22] H. Tsuei, K. Oliphant, D. Japikse, in: The Validation of Rapid CFD Modeling for Turbomachinery, IMechE Conference, London, 1999. [23] B.E. Launder, D.B. Spalding, The numerical computation of turbulent ow, Computational Mathematics in Applied Mechanical Engineering 3 (1974) 269289. [24] C. Peiderer, Die Kreiselpumpen fur Flussigkeiten und Gase, Springer-Verlag, Berlin, Gottingen, Heidelberg, 1961. [25] R.C. Dean. On the unresolved uid dynamics of the centrifugal compressor, Advanced Centrifugal Compressor, ASME Gas Turbine Division, 1971.

You might also like