Sie sind auf Seite 1von 13

Engineering Failure Analysis 13 (2006) 13381350

www.elsevier.com/locate/engfailanal

Investigation of the failure of the L-0 blades


Z. Mazur *, A. Hernandez-Rossette, R. Garca-Illescas
Instituto de Investigaciones Electricas, Av. Reforma 113, Col. Palmira, 62490 Cuernavaca, Morelos, Mexico
Received 5 September 2005; accepted 24 October 2005
Available online 14 February 2006

Abstract
A last stage (L-0) turbine blades failure was experienced at the 110 MW geothermal unit after 1 year of operation period. This unit has two tandem-compound intermediate/low-pressure turbines (turbine A and turbine B) with 23 in./
3600 rpm last-stage blades. There were exible blades continuously coupled 360 degrees around the row by loose cover
segment at the tip and loose sleeve and lug at the mid-span (pre-twist design). The failed blades were in the L-0 row of
the LP turbine B connected to the generator. The visual examination indicated that the group of 12 L-0 blades of rotor
B on the generator side was bent and another group of 5 blades at 140 degrees from the rst damaged group was also bent.
The cover segments were spread out from the damaged blades and had cracks. Laboratory evaluation of the cracking in
the cover segments indicates the failure mechanism to be high cycle fatigue (HCF), initiating at the cover segment holes
outer llet radius. The L-0 blades failure investigation was carried out. The investigation included a metallographic analysis of the cracked cover segments and bent blades, Finite Element Method (FEM) stress and natural frequency analysis
(of blades/cover segments), fracture mechanics and crack propagation analysis. This paper provides an overview of the L-0
blades failure investigation, which led to the identication of the blades vibrations within the range 250588 Hz induced
due to unstable ow excitation (stall utter) as the primary contribution to the observed failure.
2005 Elsevier Ltd. All rights reserved.
Keywords: Failure analysis; High cycle fatigue; Steam turbine failures; L-0 blade failure; Metallurgical examination

1. Introduction
The last stage blade (L-0 blade) is one of the most important contributors to the performance and reliability
of the steam turbine. With the last stage blades typically producing 10% of the total unit output, and up to
15% in some combined-cycle applications, improvements in last stage eciency can signicantly impact the
output of the total unit. Retrotting an older last stage design with a modern diaphragm and last stage blade
can typically improve heat rate and.output by up to 1% [1].
Life extension against erosion and reduction of vibration stress are important for improving the reliability
and maintainability of steam turbines last stage blades. The use of a continuously-coupled connection structure instead of the grouped blades has proved to signicantly reduce the vibration stresses and is currently
*

Corresponding author. Tel.: +52 777 3623811; fax: +52 777 3623834.
E-mail address: mazur@iie.org.mx (Z. Mazur).

1350-6307/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfailanal.2005.10.018

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

1339

implemented in many retrots of steam turbines [15]. The blades are continuously coupled 360 degrees
around the row by an integral or loose cover segment at the tip and loose sleeve and lug at the mid-span. This
provides excellent damping and reduces the dynamic response levels of the blades [1].
Modern turbomachines operate under very complex regimes where a mixture of subsonic, transonic and
supersonic regions coexists. The trend for improved large steam turbine design towards higher aerodynamic
blade loading and smaller physical size attracts much attention to the aeroelastic behavior of blades in turbines. Flow-induced blade oscillations (utter) of the turbine can lead to fatigue failures of a construction
and so they represent an important problem of reliability, safety, and operating cost. Aeroelasticity phenomena are characterized by the interaction of uid and structural domains. The cascade utter is characterized by
aerodynamic interaction among oscillating blades in the blade row. Its importance can be understood from the
fact that the unsteady aerodynamic force on blades is heavily dependent on the interblade phase angle. From
this standpoint neighbouring blade rows, e.g. a neighbouring rotor or stator will have a considerable inuence
on the unsteady aerodynamic force because blade rows are closely placed in actual turbomachines. It is known
that the ow-induced blades oscillations are pronounced mainly at the zone of the 90100% of the blade
length. Also aeroelasticity phenomena can occur while turbine is operating at low load/low vacuum. During
turbine operation at a low load/low vacuum, the L-0 blade vibration stresses are increasing abruptly. This
increase (peak) of the blade vibration is induced by unstable ow with oscillation of a shock wave near the
throat of the blade tip passage [1,2]. Operation at a low load like initial load-hold and other conditions shall
be determined by considering such phenomena.
In many cases by using either tip or mid-span shroud (cover) design, the blade structural damping can be
increased enough to prevent blade utter. However, the shrouded rotor blade design will cause the blade mode
shapes to be complex, and in some cases both bending and torsion mode components can be present at the
same time in a single mode [6]. Unfortunately, most existing analyses deal with uncoupled simple mode shape
models, which cannot be used for the combined bending and torsion system mode analysis, and judgment
based on past experience and experiments must be used to determine the acceptability of a shrouded rotor
blade design [710]. On the other hand, prediction of the forced response of shrouded disc assemblies is still
a challenging engineering task because of unknown excitation loads and friction damping eects [11].
Recently, a remarkable progress in transient ow calculations allowed the prediction of more realistic excitation forces acting on the rotating blades [1217]. Nevertheless, due to some current uncertainties in transient
ow calculation and forced response of shrouded disc assemblies, as was reported herein, some cases of steam
turbine low pressure blades failures with a continuously-coupled connection structure have been recorded.
This paper provides an overview such a case of the L-0 blades failure investigation due to ow excitation
(utter) as the primary contribution to the observed failure.
2. Background
The blade under evaluation was the 23 in./3600 rpm last stage blade (L-0) of a 110 MW geothermal turbine
which consists of two tandem-compound intermediate/low-pressure turbines with steam condition 1.1 MPag/
182 C/84 kg/s. During the unit last overhaul (January 2004), the original grouped rigid L-0 blades were
replaced by exible blades continuously coupled 360 degrees around the row by loose cover segment at the
tip and loose sleeve and lug at the mid-span. The evaluation of these failed new installed blades was carried
out after 1 year of blade operation period in base load mode. The blade is made of AISI 410 stainless steel.
During unit normal operation period the water feeding pump was tripped due to a mechanical problem and
as a result the unit vacuum was dropped from 648 to 570 mm Hg and the load from nominal to 72 MW (65%
of load) at an interval of 42 s from the last stable point. After the next 26 s the unit was tripped due to a condenser low vacuum condition. After the turbine was restarted there were measured high vibrations, which
forced the unit to be shut down-to carry out an inspection and related maintenance. The turbine visual examination revealed that the group of 12 L-0 blades from the generator side of rotor B, connected to the generator,
was bent in the direction opposite to the rotor rotation and that another group of 4 blades of the same row at
140 degrees from the rst damaged group was also bent, as is shown in Figs. 1 and 2, respectively.
Fig. 3 shows a general view of the blade cover segments spread from the damaged blades. They present
damage in the form of bending, rubbing, loss of material and cracks. The damaged loose lashing sleeves which

1340

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

Fig. 1. The group of 12 blades, L-0 row from the generator side bended at the tip.

Fig. 2. The group of 4 blades, L-0 row from the generator side bended at the tip.

Fig. 3. Damaged loose cover segments.

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

1341

couple the blades in the mid-span are shown in Fig. 4. The sleeves were separated from the damaged blades,
crushed and fractured in separate parts.
Fig. 5 shows the detail of the damage of cover segment No. 19; heavy deformation, rubbing, fracture and
separation of the pieces of material is apparent. In Fig. 6 is shown the crack initiation on cover segment No.
12. The crack is localized on the outer let radius of the wall of the cover segment hole. There were more cover
segments with similar cracks.

Fig. 4. Damaged loose lashing sleeves.

Fig. 5. Detail of the damage of the cover segment No.19.

1342

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

Fig. 6. Crack initiation on the cover segment No. 12.

3. Metallurgical investigation of the L-0 blade


The metallurgical investigation of the failed L-0 blade was carried out, and included metallography, SEM
(scanning electronic microscopy) fractography and chemical analysis. The microstructure of the blade airfoil
(bended tip zone) is shown in Fig. 7. The microstructure consists of tempered homogenous martenzite, free of
failures, typical for forged stainless steel according to specication AISI 410.
Fig. 8 shows the microstructure corresponding to cover segment number 12 made of Titanium Ti6Al4V.
The microstructure represents bimodal distribution of volumetric phase a prime (60% Vol.) within a laminar matrix of phase a + b (40% Vol.). It is a typical microstructure of Titanium alloy Ti6Al4V after
solution treatment and aging. The grain size is very small (ne grains) and measure 20 lm approximately.
Also, a lot of small transgranular micro cracks can be seen as indicated by the arrows in Fig. 8b. The transgranular crack propagation is typical of a fatigue failure mechanism.
Fractography evaluation was carried out on the exposed crack surface of cover segment No. 12 (see Fig. 6)
using scanning electronic microscopy (SEM) to determine the origin of the fracture. Fig. 9 shows the dierent

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

Fig. 7. The microstructure of the L-0 blade airfoil made of AISI 410 stainless steel (bended tip).

Fig. 8. Microstructure corresponding to cover segment number 12 made of Titanium Ti6Al4V.

Fig. 9. Fracture initiation and propagation zones on cover segment No. 12.

1343

1344

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

zones of the fracture propagation surface of cover segment No. 12. The presence of striations (fracture sliding
planes) which are characteristic for fatigue mechanism of fracture propagation, are noticeable.
Each striation represents one cycle of fatigue and by measuring the distance between striations it is possible
to determine the velocity of crack (fatigue) propagation. Average inter-striation distance was 3.73 lm in the
fracture propagation zone and 1.701.97 lm in the fracture initiation zone. On the fracture surface no beach
marks were found. The presence of beach marks on the fracture surface commonly indicates that more events
participated in fatigue propagation. The beach marks divide the fracture surface in zones of dierent roughness, which correspond to dierent events of fatigue. The absence of beach marks on the fracture surface of
cover segment means that only one event participated in fatigue propagation.
The chemical analysis of the deposits present on the fracture surface of cover segment No. 19 revealed some
quantity of Ni, Cu, Co and S. The existence of signicant clearances between the cover segment hole and the
blade tenon facilitate accumulation of the deposits (oxides) which come from other sections of the turbine and
can promote corrosion.
4. Blade stress analysis
Using the Finite Element Method (FEM), centrifugal stresses, deformation and natural frequency of the
individual blade and coupled blades at 3600 rpm were calculated. The calculation also included the determination of stresses at the cover segment. The results were evaluated considering the maximum stresses developed at the blade and cover segment, maximum deformation and possible resonances. The rst four modal
forms of vibration of the individual L-0 blade are shown in Fig. 10.
The maximum deformation of the individual blade was 6.7 mm and occured at the tip of the blade, as is
shown in Fig. 11.
Centrifugal stress distribution at the individual blade at 3600 rpm is shown in Fig. 12, The maximum stress,
451.7 MPa occurred at the mid-span of the airfoil, close to the lashing sleeves at the blade pressure surface.
The natural frequency of coupled blades is represented in Fig. 13. The rst natural frequency is 115.28 Hz,
the second is 195.52 Hz, the third is 357.17 Hz and the fourth is 358.29 Hz. It can be seen that the third and
fourth blade natural frequencies are very similar. Further, that the rst and second frequencies are larger than
those of the individual blade, whereas the third and fourth frequencies are smaller, respectively.

Fig. 10. Vibration modes of the individual blade at 3600 rpm.

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

Fig. 11. Individual blade maximum deformation due to centrifugal force.

Fig. 12. Centrifugal stress distribution at the individual blade at 3600 rpm.

First mode

Second mode

Third mode

Fourth mode

Fig. 13. The natural frequency of continuously coupled blades.

1345

1346

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

The reported natural frequencies of the blades (individual and coupled blades) conrm that no mechanical
resonances of the blades structure exist.
The maximum deformation of the continuously coupled blades was 2 mm and corresponds to the third
mode of vibration (Fig. 13). As was mentioned before a maximum deformation of 6.7 mm was registered
for the free individual blade.
The centrifugal stress distribution at the blade airfoil pressure surface is shown in Fig. 14. The maximum
stresses, of 391.5 MPa, were registered at the airfoil mid-span below the lashing sleeves and were much smaller
than in the case of free individual blade (451.7 MPa). In this view (Fig. 14) cover segments and lashing sleeves
were eliminated to see the whole blades. As may be evaluated from Fig. 15a, the maximum stress level,
510 MPa, was encountered at the surface of the cover segment in the contact zone between the blade airfoil
and the cover segment. The stress level at the hole of the cover segment in the contact zone between the blade
tenon and the cover segment (fracture zone) was lower; 250 MPa, as is indicated by arrow in Fig. 15b. The
stresses at the lashing sleeves were very low and for this reason they are not shown here.
Considering a yield stress of 574 MPa for a blade made of AISI 410 stainless steel, a yield stress of 925 MPa
for the cover segment material (Titanium Ti6Al4V), and the maximum calculated stresses presented previously, the security factor for individual free standing blade is 1.27, for coupled blades it is 1.47 and for the
cover segment it is 1.81.

Fig. 14. Centrifugal stress distribution at continuously coupled blades.

Fig. 15. Stress distribution at the cover segment.

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

1347

5. Fracture propagation analysis of the cover segment


A fracture propagation analysis of the cover segment was carried out, based on the metallographic investigation ndings and rules of fracture mechanics. Stable fatigue fracture propagation is governed by Paris Law
[18] according to
da
m
CDK mm=cycle
dN

da
where dN
is the velocity of fracture propagation; da the crack size increment during one cycle of fatigue; dN
the number of fatigue cycles; C and m the Empiric constants; DK the fatigue stress intensity.
The fatigue striations distances on the cover segment fracture surface correspond to the crack size increment, da, during one cycle of fatigue. These distances were measured and its values fall within the range
1.7 lm at the fracture initiation zone to 4 lm at the fatigue fracture end zone. It was observed that the predominant inter-striation distance was between 3 and 4 lm. To successfully separate the cover segment completely from the blade, the maximum crack size should be more or less the same as a depth of the cover
segment hole in which the blade tenon is installed (Da = 7 mm). The total time of the fatigue event was
26 s, which correspond to unit low load, low vacuum operation (72 MW, 570 mm Hg). This time includes
the periods of fracture initiation and propagation, the time required for the deformation and spread o
of the cover segment, and also the time required for the blade airfoils to deform. Considering, arbitrarily that
the time for fracture propagation was a fraction of the total fatigue event time of 7 s, the possible blade excitation frequency that could lead to the cover segment fatigue failure was calculated from

Da=Dt
da=dN

HZ;

where f is the blade excitation frequency; Da = 7 mm the depth of the cover segment hole; Dt = 7 s the time of
the fracture propagation; da = 1.7, 3 and 4 lm.
Using a rearranged form of the Paris Low the stress intensity factor, DK, was determined from
r
p
m da 1
DK
MPa m;
3
dN C
where C = 2.66e-12 and m = 4.23 determined using nCode [18].
For three options of striations distance, da = 1.7, 3 andp
4 lm, the corresponding stress intensity factors
p and
associated excitation frequencies
were
DK
=
23.57
MPa
m
and
f
=
588.23
Hz,
DK
=
26.96
MPa
m and
p
f = 333.3 Hz, DK = 28.86 MPa m and f = 250 Hz, respectively. Further, considering these frequencies
and the same options of striation distance, the length of fracture was calculated using
da
Da
 f  t mm.
4
dN
It was found that for frequency f = 333.3 Hz and striations distance da = 3 lm the fracture length was
Da = 7.5 mm, which is very close to the real cover hole depth (Da = 7 mm) considered in this analysis. This
frequency (333.3 Hz) is very close to the natural frequencies of the third and fourth modes of vibration of
the coupled blades (357.2 and 358.29 Hz, respectively). Because the striations distance 3 lm was found predominantly on the cover segment fracture surface, it can be concluded that probably the third torsional mode
X, which is typically related to the turbine operation with low load low vacuum, may be responsible for the
cover segment fatigue failure and its separation from the blades.
6. Discussion of the results
The results of the metallographic examination of the failed L-0 blade and cover segment indicate that the
crack initiation and propagation on the cover segment was driven by a high cycle fatigue mechanism. Striations characteristic to high cycle fatigue were found throughout the whole fracture surface of the cover segment. Also, whole fracture surface presented small transgranular cracks, typically related to fatigue
fracture mechanisms.

1348

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

Beach marks were not encountered on the cover segment fracture surface. The presence and number of
beach marks commonly indicate how many fatigue events participate in fatigue propagation; the beach
marks divide the fracture surface in zones of dierent roughness that correspond to dierent fatigue
events. The lack of beach marks on the fracture surface means that only one fatigue event participated
in fracture propagation. The fatigue striations distance corresponds to the velocity of stable fracture propagation da/dN.
Considering the L-0 blades real operation period (1 year), it may be concluded that the mechanical resonance of the blades does not contribute to blade failure. This conclusion also indicates that fatigue failure
of the blades/cover segments was not originated during continuous operation under vibration stresses, but
during transition events. If the fatigue initiation and propagation were during continuous operation under
resonance vibratory stresses, the blade failure would occur practically immediately (after a few hours of
operation). Analyzing the units operation history since the date in which the L-0 blades were installed (retrot), only one period of unit operation with low load/low vacuum lasting 26 s approximately, was
detected; this is congruent with metallographic ndings on the cover segment fracture surface (lack of
beach marks).
According to [2,19], the steam turbine operation with low load/low vacuum is inducing L-0 blade excitation
(vibration) by unstable ow developing high vibratory stresses. Fig. 16 shows steam ow stream lines distribution at the L-0 stage during low load/low vacuum operation [19].
Due to reduced mass ow, the steam conditions are variable along the steam path; there are zones of different pressure, radial ows, counter ows, ow recirculation (ow instabilities). These operation conditions
generate blade excitation forces (torsional vibrations X mode), which can lead to blade failures. Unit operation with reduced mass ow is also causing a reduction of the ow velocity in the same degree. This results in
changes of the blade entry ow incidence angle (change of stage velocity triangle); the ow is entering into the
L-0 blades with negative incidence angle (in this case it was 3336 approximately) striking the suction surface
of the blade airfoil and exciting the blades (stall utter). The pressure uctuation, ow recirculation and counterows, in conjunction with the negative incidence angle ow striking on the blades, developed excessive
vibratory stresses causing fatigue fracture and spread out of cover segments. In turn, it caused blade structural
loosening and drastically changed the blade vibration damping characteristics; the blades operate as it they
were free-standing. Furthermore, the forces developed by the steam ow and the vibratory stresses nally
caused the deformation of the blades. The deformation of the free-standing blades was facilitated by the moderate stress safety factor used to design the blades.

Fig. 16. Steam ow stream lines distribution at the L-0 stage during low load/low vacuum operation [19].

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

1349

7. Conclusions
On the basis of the analysis of the results of the L-0 blade/cover segment metallographic examination, unit
operational parameters, blade natural frequency and fracture mechanics it may be concluded that the L-0 blade
failure was originated at cover segments which were fractured and separated from the blades, causing loosening
of the blades and changing drastically their vibration damping characteristics. The cover segments fracture initiation and propagation was driven by a high cycle fatigue mechanism and was probably due to the combined
eect of the unit operating in low load/low vacuum conditions, resulting in transient excitation of the blades.
Considering the real L-0 blades operation period (1 year) it may be concluded that the dynamic characteristics of the blades (mechanical resonances) did not contribute to blade failure. This means that the fatigue
fracture of the cover segments was originated probably during transient events and not during continuous
(stable) operation under vibratory stresses.
Analyzing the unit operational history and the cover segment fracture surface, only one transition event,
e.g. unit trip due to operation with low load low vacuum, was found, which was associated with fatigue striations on the cover segments fracture surface with lack of a beach marks.
The turbine operation with low load and low vacuum typically results in steam ow instabilities-pressure
uctuation and ow recirculation. Also, due to reduced steam ow/steam velocity, the ow is entering into the
stage with negative incidence angle striking the suction surface of the blade airfoil. As a result, high vibratory
stresses were developed in the blades structure, causing fatigue failure of the cover segments and their spread
out from the blades, loosening the blades structure damping vibration system and a consequently the blades
operated as free-standing.
The force of the steam ow with negative incidence angle, together with vibratory stresses, generate excessive stresses and deformations in the free-standing exible blades, as well as partial blockage of the steam passage, in which ow recirculation and counter ows also participate. The free standing blades deformation was
facilitated by the moderate stress safety factor applied at the design of the blades.
References
[1] Coer JI et al. Advances in steam path technology. GE Power Systems, GER-3713E, New York; 1996.
[2] Suzuki T et al. Recent upgrading and life extension technologies for existing steam turbines. ASME Power, April 57, 2005, Chicago.
p. 57782.
[3] Ortolano RJ. Recent case histories in the inspection, modication, and repair of steam turbine blading. In: Proceedings, international
joint power generation conference, ASME PWR, vol. 13; 1991. p. 14754.
[4] Kubiak JA, Carnero A, Mazur Z. Modication of L1 moving blades. In: Proceedings, international joint power generation
conference, vol. 3, ASME PWR, vol. 28; 1995. p. 28994.
[5] Puri A, Lam T. Cumberland last stage blade failure investigation. In: Proceedings, international joint power generation conference,
vol. 3, ASME PWR, vol. 28; 1995. p. 40313.
[6] Hsiao-Wei D. Chiang, Chi-Chih Chen, Chih-Neng Hsu. An investigation of turbomachinery shrouded rotor blade utter. In:
Proceedings of ASME Turbo Expo 2003, Power for Land, Sea, and Air, June 1619, 2003, Atlanta, Georgia, USA.
[7] Kielb JJ, Abhari RS. Experimental study of aerodynamic and structural damping in a full-scale rotating turbine, transaction of the
ASME. J Eng Gas Turbines Power 2003;125:10212.
[8] Grin JK, Labelle RF. A rational method for optimizing shrouded damping, ASME Paper 96-GT-402.
[9] Sextro W. The calculation of the forced response of shrouded blades with friction contacts and its experimental verication. ASME
Paper No. 2000-GT-540 2000.
[10] Sextro W. Dynamical contact problems with friction-models, methods, experiments and applications, habilitationsschrift. Lecture
notes in applied mechanics, vol. 3. Heidelberg: Springer Verlag; 2002.
[11] Szwedowicz J et al. On forced vibration of shrouded turbine blades. In: Proceedings of ASME Turbo Expo 2003, Power for Land,
Sea, and Air, June 1619, 2003, Atlanta, Georgia, USA.
[12] Filsinger D et al. Approach to unidirectional coupled CFD-FEM analysis of axial turbocharger turbine blades, transaction of the
ASME. J Turbomachinery 2002;124:12531.
[13] Kielb R et al. Blade excitation by aerodynamic instabilities-a compressor blade study. In: Proceedings of ASME Turbo Expo 2003,
Power for Land, Sea, and Air, June 1619, 2003, Atlanta, Georgia, USA.
[14] Kielb R et al. Flutter of low pressure turbine blades with cyclic symmetric modes-a preliminary design method. In: Proceedings of
ASME Turbo Expo 2003, Power for Land, Sea, and Air, June 1619, 2003, Atlanta, Georgia, USA.
[15] Silkowski PD et al. CFD Investigation of aeromechanics. In: Proceedings of ASME Turbo Expo 2001, Power for Land, Sea, and Air,
June 47, 2001, New Orleans, Louisiana, USA.

1350

Z. Mazur et al. / Engineering Failure Analysis 13 (2006) 13381350

[16] Panowsky J, Kielb RE. A design method to prevent low pressure turbine blade utter. J Eng Gas Turbines Power 2000;122:8998.
[17] Vogt DM, Fransson TH. A new turbine cascade for aeromechanical testing. The 16th symposium on measuring techniques in
transonic and supersonic ow in cascades and turbomachines, Cambridge, September 2002, UK.
[18] Program nCode, International Limited, Version 1; 2001.
[19] Troyanowskij BM, Filipow GA, Bulkin AE. Parovye y Gazovye Turbiny Atomnych Elektrostancyj. Moscow: Energoatomizdat; 1985
[in Russian].

Das könnte Ihnen auch gefallen