You are on page 1of 54

500 Steam Turbines

Abstract
This section of the manual explains how steam turbines work and describes steam
turbine components. It also describes various control devices, auxiliary systems,
and startup and operation of a steam turbine. Rerating of steam turbines is also
discussed.

Contents Page

510 Historical Overview 500-3


520 Engineering Principles 500-4
521 Energy Conversion
522 Turbine Types and Performance Characteristics
523 Staging
524 Admission
525 Power Cycles
530 Machine Components and Materials 500-17
531 Nozzles
532 Blading
533 Rotors
534 Casings
535 Insulation
536 Steam Chest
537 Diaphragms
538 Seals
539 Bearings
540 Control 500-33
541 Governors
542 Control Strategy

Chevron Corporation 500-1 March 1996


500 Steam Turbines Driver Manual

543 Control Stage


544 Emergency Controls
550 Operation 500-43
551 Slow Roll and Warm Up
552 Critical Speeds
553 Steam Conditions
554 Shaft Seals
555 Seal Support Steam Systems
560 Instrumentation 500-49
570 Auxiliaries 500-50
571 Surface Condenser
572 Lube and Control Oil Systems
573 Piping Systems
580 Rerates 500-50
581 Single Stage Turbines
582 Multi Stage Turbines
583 Rotordynamics

March 1996 500-2 Chevron Corporation


Driver Manual 500 Steam Turbines

510 Historical Overview


The basic operating principles of the steam turbine have been known to man for
hundreds of years. The earliest steam turbine on record appears to be Hero’s
rotating sphere, Alexandria, 120 BC. Many others followed. Most early versions
were interesting demonstrations of the scientific basis; but not until the last quarter
of the 19th century did a practical machine emerge.
Working independently, and along slightly different lines, de Laval (Sweden, 1884)
and Parsons (England, 1884) both produced turbines with usable power outputs.
The de Laval turbine was a single stage, flexible shaft machine, producing about 3
HP at 26000 RPM as shown in Figure 500-1. The Parsons turbine was a double-
flow, multi stage unit, producing 10 HP at 17000 RPM as shown in Figure 500-2.

Fig. 500-1 de Laval’s Turbine, 1884. From Steam and Fig. 500-2 Parson’s Turbine, 1884. From Steam and
Gas Turbines, by Stodola and Lowenstein, Gas Turbines by Stodola and Lowenstein,
Vol. I, 1927. Vol. I, 1927.

Development proceeded slowly at first, hampered by conservative attitudes, mate-


rials, and manufacturing limitations. Nevertheless, by 1900, Parsons was producing
generator sets of 1000 kW output at 1500 RPM and had licensees worldwide.
From this point, development was so fast that by 1910, every major military nation
had largely re-equipped their navies with steam turbine power. Electric power gener-
ation for domestic and industrial consumption became commonplace in the devel-
oped world.
By about 1930, the development of the steam path and expansion efficiency was
substantially complete, and has improved only marginally since then. Considerable
room for improvement remained, however, in the areas of construction, reliability,
and materials for higher pressures and temperatures. Unit power output continued
to grow along with population and general affluence until about 1970. At that point,

Chevron Corporation 500-3 March 1996


500 Steam Turbines Driver Manual

a decline set in because of conservation concerns, the growth of cogeneration, and


the development of the gas turbine and alternative energy sources.
Steam turbines will be needed for many years, particularly in such areas as cogener-
ation bottoming cycles and process waste heat recovery.

520 Engineering Principles

521 Energy Conversion


The steam turbine converts thermodynamic energy in steam (pressure, temperature,
enthalpy) into mechanical energy (horsepower) in two phases:
• Conversion of thermal energy (enthalpy) into kinetic energy (velocity).
• Conversion of kinetic energy into work (horsepower).

Conversion of Enthalpy to Velocity


Most single stage steam turbines designed in the USA develop at least sonic
velocity in the inlet nozzle row. The first stage in most multi stage turbines
develops supersonic conditions (approximately Mach number 1.4). The relationship
between enthalpy and velocity is shown in this equation:

H 1 – H 2 = ( v 2 ⁄ 2g ) ⁄ J
(Eq. 500-1)
where:
J = mechanical heat equivalent
H1 = steam enthalpy upstream of the nozzle
H2 = steam enthalpy downstream of the nozzle
v = steam jet velocity downstream of the nozzle
Acceleration of steam occurs in nozzles. Sonic velocity, C, occurs at the throat, the
narrowest part of the nozzle, if the pressure drop is great enough. C is calculated
from this equation:

C 2 = kgZRT t
where:
Tt = temperature at the throat of the nozzle
k = specific heat ratio (1.3 for steam)
Z = gas compressibility
R = gas constant (85.83 ft lb/lb °F for steam)

March 1996 500-4 Chevron Corporation


Driver Manual 500 Steam Turbines

In practical terms, temperature is the only variable in this relationship. For the
conditions found in most steam turbines, sonic velocity lies between 1400 and 2000
fps.
Critical pressure ratio, Pc, is the required pressure ratio across a convergent nozzle
for sonic velocity:

Pc = {2/(k+1)}k/(k-1)
and
Tt/T1 = {Pc}(k-1)/k
where:
k = dry saturated steam is 1.3
T1 = absolute temperature at nozzle inlet
Tt = absolute temperature at the throat
Sonic velocity is reached in the throat. If the nozzle has a divergent section down-
stream of the throat, the flow continues to accelerate. This acceleration occurs only
if there is critical pressure at the throat and if the down-steam pressure is below
throat pressure. Without the required pressure drop, the flow actually decelerates in
the divergent section, and turbine power is very poor.
Typical single stage turbines have sonic rather than supersonic velocity steam jet
for these reasons:
• It is easier and more economical to design and manufacture a convergent
nozzle than a fully-profiled, convergent-divergent nozzle.
• Single stage turbines are often controlled by inlet throttling, again for cost
reasons. Inlet throttling requires an excess pressure drop across the turbine to
allow for the pressure drop across the throttle valve at low flow rates, while
maintaining sufficient pressure drop across the nozzles to achieve sonic
velocity. With supersonic nozzle design (convergent-divergent), any pressure
drop below design causes velocity deceleration in the divergent section of the
nozzles and reduces conversion of energy to velocity. Thus, a convergent diver-
gent nozzle is not able to operate efficiently enough at off-design conditions to
make it a worthwhile feature on a throttle-controlled turbine.
Most multi stage turbines have supersonic steam jet velocity (typically 1500-3000
fps) in the first stage. Control does not usually depend on variable pressure drop
across a throttle valve. Therefore, the nozzles face only one set of conditions and
are not exposed to excess pressure drop or the shock buffeting that goes with it. The
remaining stages are usually subsonic.

Conversion of Velocity to Power


Exactly how this function is performed depends on blade and nozzle design. (For
more information, see the section on impulse/reaction blading in Section 522.) The
basic premise is the same for both impulse and reaction design: change of velocity
creates a force which, if acting on a moving blade, produces power.

Chevron Corporation 500-5 March 1996


500 Steam Turbines Driver Manual

In Figure 500-3, using the impulse design for illustration, the nozzle outlet speed is
C1, the stage outlet speed is C2, and the blade is moving at U ft/sec. Let us assume
that the steam mass flow is m lb/sec. The thrust on the blade, F, in the direction of
blade movement, is calculated from the product of the mass flow, m, and the
change of velocity.
F = m (C1 Cos ∝1 - C2 Cos ∝2)
Power is simply force times blade speed:
HP = U × F
Maximum power and efficiency occur when C2 is minimum. Impulse stages are
most efficient when:
Blade speed ratio (U/C1) = (Cos ∝1)/2
Most single-row impulse turbines have optimum blade speed ratio of approximately
0.485 because the inlet angle ∝1 must be positive. The nozzle centerline is usually
10-14 degrees to the blades.
Efficiency cannot be greater than Cos2 ∝1 (about 94%). In reality, many losses must
be considered. The stage will probably not exceed 80%, because nozzle efficiency
is unlikely to be more than 90% to 92%.
Figure 500-4 shows the typical relationship between blade-speed ratio and effi-
ciency.

Fig. 500-3 Conversion of Velocity to Power Fig. 500-4 Relationship Between Blade-Speed Ratio
and Efficiency

Reaction-design turbines have similar underlying principles, but the analysis is


somewhat more complex. This design requires that a substantial pressure drop be

March 1996 500-6 Chevron Corporation


Driver Manual 500 Steam Turbines

introduced in the moving blades. Optimum performance occurs at a different blade-


speed ratio. Efficiencies are usually up to 3% higher, but manufacture is more diffi-
cult and expensive, because more stages as well as closer clearances are required
around the blade tips.

The Condition Line


If we plot the steam condition as it proceeds through the turbine, the resulting line
is called the Condition Line. You can see it on the Mollier Chart in Figure 500-5
(point A to point B).
In some turbines, the condition line crosses the saturation line, below which the
steam becomes wet or “condenses.” The separating water can cause much damage
through erosion and corrosion. Usually, in a well-designed turbine, the water is
partially removed through strategically-placed casing drains. In practice, the
condensation does not begin until the steam enthalpy is 97% of the theoretical
“dry” enthalpy. This supercooling is due to the finite time required for the physical
process to take place. A 3% wet line drawn on the chart below the saturation line is
called the “Wilson Line,” and it is an important consideration in the design of
casing drains.

522 Turbine Types and Performance Characteristics


Turbine Types
There are three bases by which to classify steam turbines:
• Condensing and Backpressure
• Single and Multi Stage
• Reaction and Impulse
Condensing and Backpressure. In the condensing turbine, the steam is expanded
down to very low pressures, typically 2.5 to 3.5 inches of mercury absolute (about
1.5 psia). At these pressures, the steam temperature is low, about 140°F, and the
steam is wet, having begun to condense. During the expansion process, the removal
of useful heat is maximized. All remaining heat is then rejected to a cooling water
system through a surface condenser. The condensing function, which starts in the
turbine, proceeds only to about 90% quality in the last blade row. The condenser
completes the condensation process to 100% water, i.e., 0% steam quality.
The backpressure turbine rejects its unused energy at a pressure high enough to be
of use to other processes. Normally, the exhaust steam, in this case, retains some
superheat. Contrary to the popular misconception that condensing turbines are less
efficient than backpressure turbines, the expansion efficiency is virtually identical
for both. It is true, however, that condensing cycles may be less efficient in practice
because the rejected heat is not hot enough to be of value. If there were no use for
the steam at the exhaust pressure of a backpressure turbine, that cycle too would be
inefficient. Clearly, we must look at both the expansion (machine) efficiency and
cycle efficiency when doing energy audit or analysis.

Chevron Corporation 500-7 March 1996


500 Steam Turbines Driver Manual

Fig. 500-5 Mollier Chart Showing a Typical Turbine Condition Line. Courtesy of the Babcock and Wilcox Company.

March 1996 500-8 Chevron Corporation


Driver Manual 500 Steam Turbines

Single and Multi Stage. Staging refers to the number of discrete pressure reduc-
tions in the turbine.
A single stage turbine typically has one set of nozzles that drops the pressure from,
for example, 300 to 100 psi. It may have more than one row of moving blades but
only one location in the turbine where pressure energy is converted to velocity
(kinetic) energy.
A multi stage turbine, on the other hand, has several nozzle rows that reduce the
pressure. A typical pressure ratio across a stage is approximately 0.58. Thus, for six
stages the overall pressure ratio would be 0.586 or 0.038. If the inlet pressure were
1000 psia, the backpressure would be 1000 x .038 or 38 psia. Such exhaust steam
would be suitable for process heating.
In reality, the pressure ratio of all six stages would not be equal. Because of the
need to control the steam flow and distribution in the turbine, the first stage, usually
called the control stage, would have a higher pressure drop, typically 0.25 to 0.35.
See Section 543.
Alternatively, the same staging arrangement could take 50 psia inlet steam, expand
to 1.9 psia (3.8 inches Hg) and exhaust to a surface condenser (cooler). Note that
although the same blading aerodynamic design could work as efficiently on both
sets of steam conditions, the power output would be much less at the lower pres-
sure, and the loads on the blading would be substantially different.
Impulse and Reaction. The way in which the pressure drop is distributed between
the rotating and stationary blades distinguishes impulse from reaction turbines. The
two have very different requirements in blade geometry and assembly tolerances.
Impulse is the force created on an object, such as a flat plate or blade, when it slows
a jet of fluid.
The impulse turbine stage drops the available pressure in a nozzle to create
maximum velocity. The jet is then played onto the moving blades to convert the
velocity (kinetic energy) into work. There is no pressure drop in the moving blades;
but in the ideal situation, the outlet velocity from the moving blades is minimal. See
Section 521 above.
Reaction is the thrust force resulting from the creation of a high velocity jet. A
good example is a fire hose pushing against the direction of the water flow. This
thrust is the reaction force.
A reaction turbine, like the impulse turbine, has fixed nozzles. The steam jet plays
on the moving blades and creates a force by virtue of impulse through momentum
exchange. In addition to the impulse force on the blade, pressure drops within the
moving blade row, and the fluid is re-accelerated. The force arising on the moving
blades as a result of the velocity change within the blade passage is the reaction
force.
The total work on the moving blades is the sum of the reaction and the impulse
forces multiplied by the blade speed.

Chevron Corporation 500-9 March 1996


500 Steam Turbines Driver Manual

The ratio of the reaction work to the total work is called “degree of reaction” and is
typically limited to 50%.
Steam velocities tend to be lower in the reaction turbine, although total work done
is theoretically identical. In reality, the reaction turbine is up to 3% more efficient.
However, it requires closer clearances around the blade tips, is more sensitive to
variations in blade section geometry, and typically needs more stages for a given
pressure ratio.
In practice, reaction turbines are more efficient when the physical dimensions are
large. Impulse turbines, on the other hand, are more efficient when small and are
traditionally considered to be more reliable.
Because they face high energy costs, European makers have concentrated on devel-
oping the reaction type. U.S. makers, facing lower energy costs, have concentrated
on the more reliable impulse types.
Although much of the original development was in impulse types, many current
machines combine impulse and reaction blading in the same casing for the best of
both worlds. The first stage of all multi stage units is invariably the impulse type.

Performance Characteristics
Torque/Speed Relationship. At maximum steam flow, all steam turbines show a
similar torque characteristic. The relationship is linear; as the speed reduces, torque
increases. See Figure 500-6.

Fig. 500-6 Typical Torque/Speed Relationship for Steam Turbines

March 1996 500-10 Chevron Corporation


Driver Manual 500 Steam Turbines

Speed/Efficiency Relationship. At any steam flow, efficiency varies with


blade/speed ratio as follows:

Efficiency = A(u/c)(Bcos∝-u/c)
where:
A, B, and ∝ = constants depending on the blade and nozzle geometry
u/c = ratio of blade speed to steam jet speed
Figure 500-4 shows a typical plot of efficiency versus speed for a single stage,
single blade row turbine.

523 Staging
Velocity Staging
Rows of moving blades that do not exhibit pressure drop are known as velocity
stages. The combination of many such rows is known as velocity compounding.
Figure 500-7 (left hand side) shows a section of nozzles for a pressure stage.
Curtis Stage. A subset of velocity compound staging is called the Curtis stage,
named after the man who took out early patents. It combines one pressure-drop
nozzle row followed by two velocity stages (see Figure 500-7). This design was
typical of many early American turbines, and remains so today, particularly for
such manufacturers as GE and Worthington (Dresser). In fact, almost all multi stage
turbines made today worldwide have Curtis design first stages. Typically the pres-
sure drop across the stage is large (ratio 0.25 to 0.4).

Pressure Staging
A pressure stage, in either impulse or reaction design, comprises a single row of
nozzles exchanging pressure for velocity in one operation, followed by a row of
moving blades. In the industry, this arrangement is conventionally known as a de
Laval turbine.
The “Rateau turbine” has two or more pressure stages in series (an arrangement
also known as pressure compounding or Rateau staging). Figure 500-7 shows the
velocity and pressure distribution as the steam flows through successive stages.
Pressure staging is on the right (downstream) side of the figure.

524 Admission
A steam chest admits steam into the turbine. The steam is then directed to the first
stage by means of an arc of nozzles. See Figure 500-8.
The nozzles are divided into sections or blocks. The complete arc of nozzles rarely
covers more than 90 to 180 degrees of total circumference, as illustrated in
Figure 500-9. This arrangement, known as partial admission, is an important part of
the power and speed control system.

Chevron Corporation 500-11 March 1996


500 Steam Turbines Driver Manual

Fig. 500-7 Steam Flow Through Turbine Stages. Courtesy of The Elliott Company.

March 1996 500-12 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-8 Nozzle Block Installed

Fig. 500-9 Partial Admission With Multiple Nozzle Valves

Full admission, in which the inlet steam is admitted to the whole 360 degrees, is
rare. More commonly seen in combustion gas turbines, it would be desirable in

Chevron Corporation 500-13 March 1996


500 Steam Turbines Driver Manual

steam turbines only at very high temperatures (above 1100°F). The purpose of full
admission in this context is to achieve thermal symmetry and avoid casing distor-
tion and warping. This type of steam turbine is not commercially available,
although it has been considered in energy-efficient experimental designs.
Note that downstream of several pressure-compounded stages, the flow that started
out as partial admission distributes evenly around the circumference, becoming,
effectively, full admission.

525 Power Cycles


Rankine Cycle
All steam turbines work on some variation of the Rankine cycle. See Figure 500-10.
This cycle is practical and depends on the following factors:
• Compression of the working fluid in liquid phase, points 1 to 2.
• Heat transfer from low temperature source, boiler feedwater preheat, points 2
to 2′.
• Boiling to change phase to vapor, boiler points 2′ to 3.
• Expansion of vapor work/heat interchange, points 3 to 4.
• Condensation of residual vapor to liquid phase, condenser, points 4 to 1.
• Re-compression of liquid working fluid.
There are many variations of the Rankine cycle. These are mainly of interest in elec-
tric power generation. In a process plant, the need for flexibility usually drives the
user toward the simple cycle, although overall thermal efficiency may be lower. See
Figures 500-10, 500-11, and 500-12.

Fig. 500-10 Rankine Cycle. From Introduction to Thermodynamics: Classical and Statistical by Sonntag and Van
Wylen. 1971. Courtesy of John Wiley & Sons, Inc., New York.

March 1996 500-14 Chevron Corporation


Driver Manual 500 Steam Turbines

Figure 500-10 shows a basic Rankine cycle without superheat. It has saturated
steam at the turbine inlet. The turbine runs almost entirely in the “wet” region.
Although attractive from a boiler-design viewpoint, this design is not optimal. Not
only is the turbine susceptible to severe erosion, but removing the condensate from
the steam path to the required degree is extremely difficult.

Superheat
To surmount erosion problems and improve cycle efficiency, we commonly super-
heat the steam to several hundred degrees above the saturation temperature, as
shown in Figure 500-11. The maximum practical superheat, also called the metallur-
gical limit, is approximately 1100°F. At higher values, exotic, temperature-resistant
materials (rarely used in the industry) are required.

Fig. 500-11 Rankine Cycle With Superheat. Courtesy of the Babcock and Wilcox Company.

Also, above 1100°F, thermal distortion of the turbine structure becomes a potential
problem, and keeping the pressure containment tight is difficult. These issues can
be addressed by making the design axisymmetric (full-admission designs, forged
cylindrical or barrel type construction, etc.). Although such designs have been
successfully deployed, there are no such machines available commercially.

Reheat
Reheat is a cycle variation in which the partially-expanded steam is extracted from
the turbine steam path, passed back to the superheater, where it is raised to its orig-
inal temperature, and then returned to the turbine, where the expansion process
continues. See Figure 500-12. It is common to have two or more reheats in utility
steam turbine generator plants. Industrial process plants, however, do not usually
use reheat because the gain in efficiency is not worth the added complexity.
Regenerative Cycle. The basic premise of the regenerative cycle is to remove some
thermal energy from the part expanded working fluid and add it to the incoming
fluid prior to the addition of energy from the outside source (for example, raise the
boiler feedwater to the saturation temperature before it enters the boiler). The aim is

Chevron Corporation 500-15 March 1996


500 Steam Turbines Driver Manual

Fig. 500-12 Rankine Cycle With Reheat. From Introduction to Thermodynamics: Classical and Statistical by Sonntag
and Van Wylen. 1971. Courtesy of John Wiley & Sons, Inc., New York.

to raise the mean temperature at which the main heating process occurs, thus
improving the Carnot efficiency of the cycle.
Recuperative Cycles. Recuperation is another cycle variation in which thermal
energy from the spent fluid stream is added to the incoming fluid prior to the main
addition of energy from the outside source. Such cycles are common in gas turbine
plants, where the heat rejection temperature is of necessity very high (1000°F). In
condensing steam turbine cycles, recuperation is not of great value because of the
low heat-rejection temperature. For back pressure cycles, recuperation may occa-
sionally be of value.

Topping and Bottoming Cycles


A Rankine cycle is often integrated with process-waste heat recovery or added onto
other systems as a topping or a bottoming “cycle.”
These are not true cycles, as the term really applies to the position in the “hier-
archy” of energy usage. For example, process-waste heat is used to generate 1500
psi steam, which in turn drives a critical compressor or generates electric power, the
pass out or exhaust steam from which is used for process heating. The steam
turbine cycle, in this case, is the topping cycle.
More commonly, however, the topping cycle uses a gas turbine, and the exhaust
heat is used to generate steam that is then supplied to a steam turbine. The two
turbines can be coupled to the same generator. In this arrangement, the gas turbine
is the topping cycle and the steam turbine is the bottoming cycle.
Condensing steam turbines are quite suited to bottoming duty because of their effec-
tive use of low-temperature energy. Back-pressure turbines or gas turbines, on the
other hand, typically reject heat at several hundred degrees F.

March 1996 500-16 Chevron Corporation


Driver Manual 500 Steam Turbines

530 Machine Components and Materials

531 Nozzles
Nozzles convert pressure energy into velocity (kinetic energy). The two basic
nozzle types are convergent, which achieves sonic velocity, and convergent diver-
gent, which achieves supersonic velocity.

Sonic Nozzles
These nozzles are usually found in small, single stage, throttle-controlled turbines.
The sonic nozzle is of converging cross section only. It is usually drilled and may
be reamed to achieve a smooth and continuous reduction of diameter. The passage
is usually straight, with the outlet set at an angle of approximately 10 degrees to 14
degrees to the rotor blade row. Efficiency is no more than 90% to 92%.
Major losses are due to the oblique inlet and discharge. Pressure ratio (P2/P1) is
usually no more than approximately 0.58, and enthalpy conversion in the steam is
limited to about 75 BTU/lb.

Supersonic Nozzles
Supersonic nozzles are usually found in the control (first) stage of a multi stage
steam turbine. They can be manufactured several ways, but their form is usually a
blade row or cascade, as shown in Figure 500-13.

Fig. 500-13 First-Stage Nozzle Ring. Courtesy of The Elliott Company.

The nozzle cross-sectional area is rectangular, which better matches the moving
blade row entrance, and the flow profile is convergent divergent. See Figure 500-14.
Although a few rare supersonic nozzles are circular or straight/oblique, most are
curved with axial entry and oblique exit. Typical efficiencies are 85% to 90%.
These nozzles have a much higher enthalpy drop than subsonic types, typically 150
to 200 BTU/lb, which gives velocities of Mach 1.4 to 1.6. They are usually made of
martensitic or ferritic chrome steel and cast or fabricated into multi-blade blocks.
These blocks are welded or bolted to the outlet side of the steam chest. See
Figure 500-8.

Chevron Corporation 500-17 March 1996


500 Steam Turbines Driver Manual

532 Blading
The two categories of blade are impulse and reaction. Impulse blades are some-
times called buckets because they function much like the buckets in early water
turbines.

Impulse Blading
The passages between the blades are designed for constant velocity without expan-
sion. The passage cross-sectional area is therefore constant. See Figure 500-14.
Often, the blade has identical inlet and outlet geometry, although on occasion the
section is rotated to make better use of the available energy and velocities.

Fig. 500-14 Impulse Blading

For short blades (low ratio of root to tip diameter) the sections are not usually
tapered or twisted. See Figure 500-15 for a definition of terms.
The blades commonly have tip shrouds, which may be integral with the blade or
riveted strip steel. Sometimes the shroud is laminated with overlapping joins, giving
the effect of a continuous band. More commonly, however, the shroud band is in
short sections (five or six blades), giving rise to the term “blade packets.”
The primary purpose of the shroud is to dampen blade vibrations and, secondarily,
to prevent leakage and flow losses around the blade tip.

March 1996 500-18 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-15 Blade Nomenclature

Figure 500-16 shows a rotor with two different arrangements of tip shroud. The two
blade rows in the foreground have integral shrouds with a continuous steel rolled
into the tip— an excellent design feature. The three rows in the background are the
lower pressure stages and have a standard packeted arrangement with a steel strip
riveted into position.
All the blades on this rotor are impulse type, clearly identifiable by the sharp
leading edges and symmetrical inlet and outlet angles.
A variety of impulse blades are shown in Figure 500-17, each one designed for a
specific application. Note the integral shrouds and rivet tangs on the shorter blades.
Often, a strip seal (labyrinth) around the tip reduces flow losses past the blade.
Theoretically, these seals are unnecessary because there should be no pressure drop
across the blades. However, in practice there is always some resistance, and the
seals do contribute to efficiency.

Reaction Blading
Reaction blading is considerably more complex in its geometry. The passage area
between blades is variable; the blade cross section is a true aerofoil. Inlet and outlet
geometry is not symmetrical, and blades are usually tapered and twisted, especially
if they are long.
Historically, reaction blades were often free standing, with feathered tips. The feath-
ered tips were designed to reduce the impact of damage by inadvertent contact with

Chevron Corporation 500-19 March 1996


500 Steam Turbines Driver Manual

Fig. 500-16 Integral and Packeted Blade Tip Shrouds. Courtesy of Dresser Rand Company, Steam Turbine Division.

March 1996 500-20 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-17 A Variety of Impulse Blades. Courtesy of The Elliot Company.

the casing. Such contact was common, because tip clearances had to be kept small
to avoid tip bypassing under the influence of pressure drop across the moving
blade. Later designs tend to favor tip shrouds and seals in the interest of reliability.

Dampers for Blade Vibration


Long blading often features lacing wire at an intermediate blade radius. See
Figure 500-18. The purpose of the wires is to provide friction damping. The wire is
placed at the vibration antinode (the point of maximum displacement under
vibrating conditions), and the centrifugal force presses the wire radially onto the
blade and restrains movement, thus providing frictional damping. Note that the wire
and blade are not firmly fixed together. Relative movement is allowed but is
restrained by friction.
Wire damping is less popular now than in times past because a hole at the bending
antinode causes inherent weakness. (The antinode is the point of maximum bending
moment.)
Other forms of damping are occasionally used, but modern designs rely more on
careful prediction of natural frequencies and conservative stressing to avoid prob-
lems. Modern analytical methods make this the preferred approach.

Blade Roots
Every blade needs to be attached to the rotor or disc. This fixture is a critical aspect
of most designs. There are, however, as many fixture variations as there are
suppliers, as was seen in Figure 500-17.

Chevron Corporation 500-21 March 1996


500 Steam Turbines Driver Manual

Fig. 500-18 Blades With Damping Wires. Courtesy of Dresser Rand Company, Steam Turbine Division.

March 1996 500-22 Chevron Corporation


Driver Manual 500 Steam Turbines

The one common feature is a hook or dovetail. Sometimes the hook is single or T-
shaped, sometimes double. Accuracy of manufacture and fit up is paramount if the
load is to be shared between two hooks.
To save on costs, blades are usually assembled onto the disc in a tangential direc-
tion, with a special locking root to close the radial entry gap.
Axial-entry, fir tree roots are preferred for high-speed or highly-loaded blades, even
though they are much more expensive. Tight tolerances are still required, but they
can be attended to individually and thus more easily.

Materials of Construction
Blade material is usually 400 series chrome steel, such as AISI 416 or 422, even
though it is somewhat susceptible to pitting corrosion or chloride stress corrosion.
Duplex alloys, used occasionally, are more expensive alternatives, but probably not
much better.
Titanium blading is rare. Most manufacturers in the steam turbine industry are unfa-
miliar with the required procedures for design and manufacture of titanium blades.

Blade Stress and Vibration


Blade stress is conventionally categorized as either steady state or vibratory.
Steady state stress is the stress resulting from rotation (i.e., centrifugally induced).
It is a tensile stress and is usually calculated using finite element methods.
Alternating or vibratory stress is usually seen as the stress arising from steam (or
gas) bending. Obviously, there is some force on the blade as a result of the steam
playing onto it. As a conservative but practical approach, we assume that the load
reaches a peak and then decays to zero as the blade passes a nozzle passage.
The critical areas are the fillet between the blade shank and the platform blade root
attachment detail. See Figure 500-15.
Typically, there are 50 to 100 blades or nozzles in a row, and turbines run at 1800 to
15000 RPM. Clearly, fatigue must be carefully addressed in the design stages
because of the many millions of stress cycles. As it takes only a few minutes to
reach 107 cycles, stress must be kept below the endurance limit.
The Goodman diagram is an essential tool in this work. Figure 500-19 shows a
typical diagram with sample values plotted.
Note that as the steady state stress reduces, the permissible alternating stress
increases. Note also that the calculated stresses must include the effects of stress
concentrations caused by root fillet and firtree radii. Typical blade root fillet stress
concentration factors in use today, for example, are 1.3 to 1.5.
A generous radius and a good surface finish are essential. Scratches and dings
caused by foreign-object damage or poor handling can cause damage quickly.
Large low-pressure blades such as those used in the last stage in a condensing
turbine are particularly vulnerable.

Chevron Corporation 500-23 March 1996


500 Steam Turbines Driver Manual

Fig. 500-19 Goodman Diagram

The critical stress location is usually at the blade root fillet, or in the first hook of
the root itself. The alternating component of stress at this location, when plotted on
the Goodman diagram, should be considerably less than the permitted maximum.
The ratio of the actual to the permitted stress is called the “Goodman Factor”. A
factor of 2 is marginal, but a factor of 5 is most desirable.

Blade Vibration
Resonance resulting from the alternating stresses must also be considered. A
screening process is used to check for interference between the frequency of the
alternating stress (the exciter) and the mechanical natural frequency of the indi-
vidual blades. The Campbell diagram represents the results graphically. See
Figure 500-20.
The purpose of the diagram is to highlight any situation where the blade natural
frequency coincides with an exciter within the normal steady operating speed
range. In this situation, resonance can set up and cause blade failure. Under reso-
nance conditions, stress is multiplied many times, by a factor dependent mainly on
damping. Without damping, the amplification factor (Q) may be up to 100.
Clearly, resonance must be avoided if possible, preferably by careful blade selec-
tion or introduction of damping features, such as lacing wires, tip shroud packeting,
etc.
The example in Figure 500-20 is interesting from a design point of view. The inter-
ference between the first lateral blade mode and the nozzle passing frequency
occurs just below the operating-speed range. This situation would be barely accept-
able. If the turbine were to stop just short of the normal speed range during accelera-
tion, it could be disastrous for the blade row in question, unless the blade is very
robust, with Goodman factor at 20 or more.
In the same example, the interference between the blade torsional and twice nozzle
passing frequency is in the operating-speed band, so a closer look is justified. If the
blade in question is a short, shrouded or packeted blade, it’s unlikely that a problem
would arise. If the blade is a long and free standing blade, additional damping is

March 1996 500-24 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-20 Campbell Diagram

justified, and a change to the blade thickness or taper ratio would be needed to raise
the response frequency above the speed range.

533 Rotors
Some rotors are built up (discs shrink-fitted to a forged shaft, as shown in
Figure 500-21) and sometimes are solid forgings (integral discs and shaft).

Fig. 500-21 Typical Built-Up, Single-Flow Rotor. Courtesy of the Elliot Company.

Chevron Corporation 500-25 March 1996


500 Steam Turbines Driver Manual

Built-up rotors usually have hot-rolled AISI 4130 shafts (4330 if over about 6 in.
maximum diameter) and forged 4140 discs.
Solid forgings are usually found on units running above about 10,000 RPM, espe-
cially if the units are multi stage.
One-piece rotor forgings are not easy to categorize; typically they are similar to
A293 class 5. Although this is a tough, moderately strong steel, hot repairs (weld
build up, for example) are difficult, especially if the yield point is over 80% of
UTS. Welding on such rotors is usually a complex procedure requiring considerable
expertise.

534 Casings
Casings perform two functions:
• Hold parts such as diaphragms and bearings in correct relative position.
• Contain the steam pressure.
When steam temperatures reach approximately 700°F or above, a double casing
design is often used, as shown in Figure 500-22. This design approach allows the
outer part of the casing to contain pressure without the attendant high temperature,
thus avoiding complications such as creep and thermal transients.
The inner casing, although exposed to the full temperature of the steam path, has
relatively modest pressures to handle. Being lowly stressed, it is better able to
handle thermal transient stresses.
During construction, casings are pressure tested to stress levels according to ASME
VIII pressure vessel code. Neither testing nor periodic re-testing, however, is
legally mandated.
The initial pressure test, especially of condensing turbines, may be staged. Blank
partitions are inserted in the casing at several locations so that the high pressure end
is tested at high pressure, mid sections are tested at intermediate pressure, etc.
If a turbine is re-rated to higher pressure, a partitioned test should be revalidated. In
spite of the major inconvenience, this requirement must never be ignored.
If the original design has been subject to a detailed finite element analysis (FEA)
and the results verified by comparing them to the strain gauge measurements in crit-
ical areas, it may be possible to qualify the casing for higher-pressure duty by rerun-
ning the FEA and comparing it to the allowable stress levels. Otherwise, a hydro
test is needed for safety reasons, if not for legal compliance.

535 Insulation
Insulation on steam turbine casings performs three functions:
• Maintains uniform temperature to avoid thermal distortion and internal
misalignment.

March 1996 500-26 Chevron Corporation


Driver Manual 500 Steam Turbines

• Protects personnel.
• Increases efficiency.

Fig. 500-22 Turbine Components. Courtesy of Borsig Turbines.

Uniform temperature is important, especially in multi stage turbines, where rapid or


uneven changes can cause contact in close-clearance areas, leaking casing joints,
and other problems.
The effect of insulation on efficiency is slight, perhaps less than 0.5%. Avoiding
thermal distortion and internal misalignment caused by uneven temperature is far
more important than achieving a small increase in efficiency. In certain cases, it
may even be better to remove insulation altogether than have parts missing or in
poor repair.

Chevron Corporation 500-27 March 1996


500 Steam Turbines Driver Manual

Personnel protection is self explanatory. The dangers of uninsulated casings are


slight compared to those of a superheated steam leak from a casing split line.

536 Steam Chest


The steam chest is a small volume between the inlet isolation valve (or trip and
throttle) and the nozzle valves where steam at header pressure accumulates for
distribution to the different nozzle sectors. See Figure 500-23.

537 Diaphragms
Diaphragms are the partitions in the casing between successive stages. See Figures
500-23 and 500-24. They carry the stationary blades or nozzles for the downstream
stage and the shaft seal between stages. Because they hold the pressure difference
between stages, they need to be strong and rigid.
Diaphragms are constructed of steel or cast iron, depending on pressure difference
and user preference.

Fig. 500-23 Turbine Casing Components. Courtesy of The Elliott Company.

538 Seals
There are three types of seals:
• Labyrinth seals
• Segmental ring seals
• Dry gas seals

Labyrinth Seals
Labyrinth seals are a close-clearance leakage seal, typically used for casing/shaft
seals and internal stage-to-stage seals. See Figure 500-25.

March 1996 500-28 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-24 Fabricated Diaphragm Designed for 100% Admission

Fig. 500-25 Labyrinth Seal. Courtesy of Elliott Company.

LABYRINTH seal showing how high and LEAKAGE flow and 1. First-stage pressure equals 300 PSIA at full load.
low sealing strips combine with stationary sealing steam 2. Leakoff to turbine stage at approximately 125 PSIA.
baffle to hinder flow of steam. The spring arrangement for
3. Leakoff to turbine stage at approximately 40 PSIA.
allows the stationary baffle to move away steam end of
from the shaft if a rub occurs. The heat condensing 4. Sealing steam at startup approximately 18 PSIA.
generated is absorbed by the stationary turbine. Conditions 5. Steam and air drawn on gland condenser at approximately 13.5 PSIA.
baffle. This protects the shaft and at right are typical
6. Small amount of atmospheric air at 14.7 PSIA drawn in.
minimizes rotor damage. values.

Chevron Corporation 500-29 March 1996


500 Steam Turbines Driver Manual

Pressure drop comes from leakage through the small annular gap (typically 0.005 to
0.010 inch). These seals are usually made of phosphor bronze or AISI 316 running
against carbon steel sleeves. This combination is moderately forgiving in a contact
situation but can cause a serious “friction whirl” if set too tight.
A friction whirl is a rotor-dynamic response to rotor-stator contact in which there is
a backwards precession relative to the forward rotation. Because the frequency of
the whirl precession is lower than the rotor speed, it shows on vibration-monitoring
display as a subsynchronous vibration, usually at a precise fractional frequency
such as 1/2 or 1/3 forward speed. The danger in such a whirl is that, being back-
wards relative to the rotation, the damping inherent in the fluid film bearings is
reduced, perhaps critically, allowing the amplitude of the whirl orbit to grow
beyond permissible limits. The large amount of available energy produced can
cause mechanical damage to the machine structure.
Other material combinations such as nickel aluminide honeycomb and 316 intro-
duce the advantage of abradability. Clearances can be relatively tight, and rotor
orbit creates the required clearance by the cutting action of the labyrinth tips. Safety
from friction whirls comes from the low friction coefficient of the seal material
combination.
On condensing turbines, a slightly positive steam pressure must be applied to the
seal to prevent air from entering the casing or condenser. Thus a supply manifold is
required (See Figures 500-26 and 500-27). During normal operation, leakage steam
from the inlet control valve stems is ideal for the purpose, but during startup, let-
down steam must be piped in from the supply header.

Carbon Ring Seals


Like labyrinths, carbon ring seals work by means of a pressure drop through small
annular gaps in a series of stages along the shaft. See Figure 500-28.
Their clearances, however, are closer than those of labyrinths for several reasons:
• Carbon has good rubbing characteristics with the shaft material.
• The ring is segmental and can adjust its diameter to fit the shaft closely. The
segments are held in place by a circumferential spring (Garter spring).
• The ring is arranged to “float” on the shaft, thus following rotor displacement.
These seals have been used for many years and can be found in most single stage
units.
Segmental carbon or graphite rings are prone to wire drawing or steam cutting if
they are left in the stationary condition with a pressure to seal. For this reason,
many standby steam turbines are slow rolled. Lately, however, there has been a
trend toward retrofitting such machines with dry gas seals. Refer to the Utilities
Manual, Appendix E, for the Company’s “Best Practices to Eliminate Turbine Slow
Rolling”

March 1996 500-30 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-26 Turbine Shaft Seal Arrangement

Dry Gas Seals


Dry gas seals are a variation of the pump mechanical seal. See Figure 500-29. The
only essential difference is that the face of the rotating seat is treated with a profile
(a rayleigh step) that develops a dynamic pressure under running conditions. The
pressure is enough to separate the faces so that there is little friction. Leakage is
minor and controlled. Under static conditions, the seal faces close up and there is
no leakage.
These seals are unsuitable for condensing turbines because they can seal only a
modest reverse pressure (about 5 psi). They can be used, but setup is very complex
and ideally requires a double seal and closely-controlled buffer gas pressure.
Such seals are available from Crane, Durametallic and BWIP.

Chevron Corporation 500-31 March 1996


500 Steam Turbines Driver Manual

Fig. 500-27 Gland Sealing and Leak-Off Equipment

Fig. 500-28 Segmental Carbon Ring Seal. Courtesy of Dresser Rand Company, Steam
Turbine Division.

March 1996 500-32 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-29 Dry Gas Seal. Courtesy of Durametallic Corporation, Kalamazoo, Michigan.

539 Bearings
See the section on bearings in the General Machinery Manual.

540 Control

541 Governors
In any steam turbine, speed and power are the two basic quantities that must be
controlled. The same governor can control both. However, the required perfor-
mance of the driven machine usually does not permit them to be controlled by the
same governor at the same time.

Chevron Corporation 500-33 March 1996


500 Steam Turbines Driver Manual

For many years, steam turbines had flyweight-type governors or a developed


version thereof (see Figure 500-30). All early governors operated in droop mode, in
which the control speed varies (droops) as the power increases (see Figure 500-31).
The simple governor droops automatically. However, many applications must have
isochronous control (constant speed regardless of control output), which requires
compensating the governor, usually hydraulically. The compensation on some
governors is adjustable, allowing the amount of droop to be controlled. Sometimes
droop is controlled to improve stability and sometimes to keep driver output at a
constant speed, such as is often required for electric power generation.

Fig. 500-30 Flyweight-Type Governor. Courtesy of Fig. 500-31 Governor Droop Phenomenon. Courtesy of
Woodward Governor Company. Woodward Governor Company.

Today, most mechanical governors are hydraulically compensated to eliminate or


control droop, and are coupled to a hydraulic servosystem to operate the steam flow
control valves.
Some direct-acting (no servo) systems are still available commercially for small
non-critical drives (See Figure 500-32). These generally do not give adequate
performance for refinery process applications.
Electronic governors and electric servosystems are becoming more common. They
can be very expensive and difficult to justify for normal refinery use but can
provide many benefits:
• Eliminate the need for high-pressure control oil systems.
• Provide better resolution of control.
• Result in a narrower dead band.
• Operate automatically (depending on setup).
• Provide compatibility with computerized plant-control strategy.
Many electronic governors are analog. Digital types are becoming more common
and are suitable for plant digital control systems (DCS) interface. Although they are

March 1996 500-34 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-32 Section of a Single-Stage Turbine and Governor System. Courtesy of The Elliott Company.

relatively expensive because it is early in their design life-cycle, competition among


suppliers is beginning to bring down the cost.

Some Typical Governors


Woodward PGPL. This governor has been used for many years on critical, vari-
able-speed drives such as centrifugal compressors that require a NEMA class D
governor. NEMA class D calls for speed control to within 0.25% of set point and,
in the event of sudden loss of load, maximum speed rise when operating at rated
conditions is 7%. Also, control speed must not vary by more than 0.5% across the
power band 0% to 100% full load.
It is an isochronous governor, controlling at a constant speed, irrespective of the
required load. See Figure 500-33.
The PGPL governor is driven through a reduction gear from the turbine non-
coupling end. It typically ranges from 70% to 110% of rated speed. The output is
by rotation of a torque arm. The torque is quite low and normally requires a servo
system.
Woodward PGG. This is a more complex governor, which can operate in isochro-
nous mode, droop speed control mode (see Figure 500-34), or load control mode, a
subset of droop speed control. It is typically used to load control a fixed-speed
machine, such as a synchronous generator connected to a utility supply grid, for
which NEMA class C is required.

Chevron Corporation 500-35 March 1996


500 Steam Turbines Driver Manual

Fig. 500-33 Isochronous Control

Fig. 500-34 Droop Control Single Machine

NEMA class C calls for speed control to within 0.25% of set point and no more
than 7% speed rise on sudden load loss. A 4% increase in steady-state control speed
is permitted for load reduction from 100% to 0% load, which can be interpreted as
4% maximum droop.
The PGG governor can operate in true isochronous mode. More often, however, it
operates in droop mode because the speed is dictated by outside influences. For
example, the frequency of the local utility supply to which the drive is connected

March 1996 500-36 Chevron Corporation


Driver Manual 500 Steam Turbines

can affect the speed. The droop is adjustable from 0% to about 10%. The effect of
changing droop would show on Figure 500-34 as a change in the slope of the
control line. Normally, droop would be set at about 3%.
Speed is selected by adjusting the speed input control. This adjustment translates
the control line up or down the page.
Load is also selected by adjusting the load-control input. This adjustment translates
the control line sideways across the page.
A typical operating scenario would be to bring the generator up to 95% of synchro-
nous speed manually, set the load at zero, set the droop at about 3%, let the auto-
matic synchronizing gear raise the speed to synchronous, and to synchronize
automatically. Finally, the load control would be manually adjusted to give the
required output. Alternatively, the final 5% speed and synchronizing can be done
manually.
Woodward TG13. The TG13 is a simple unit intended for non-critical service. It is
a NEMA class A governor with 10% fixed droop as standard, a 13% speed increase
for sudden load loss, and a 0.75% speed resolution. The 10% droop gives good
control stability with single-valve control of single stage, general-purpose steam
turbines, without the need for a servosystem.
Typically, however, the TG13 cannot be used with steam pressures above about 250
psig, depending on valve size, as the control valve loads are too high.
The TG13 has many refinery applications because it is ideally suited to single stage
turbine pump drives. It is often used in upgrades from manufacturers’ standard inte-
gral non-servo governors, such as those found on Elliott and Terry general-purpose
turbines.

Electronic Governors
Electronic governors are becoming more common. They typically have very good
resolution (0.1%), a narrow deadband (0.1%), and an excellent control range (10%
to 100%). They can be extremely fast and commonly achieve load-loss speed-rise
better than 5%.
A number of electronic governors are on the market. They can be preprogrammed
to run drives up to speed smoothly, either at a predetermined rate or in stages. Thus,
it is possible to be absolutely certain that critical speed bands are traversed properly
and promptly.
Most units rely on a toothed wheel and magnetic or proximity probes to generate
the speed signal. Some units offer redundancy by duplicating circuits. API 612
fourth edition calls for duplication of the pickups, but the better units already offer
the vastly superior two-out-of-three voting logic.
Woodward 2301. The 2301 is an analog unit intended for mechanical drives and
generators. It can perform isochronous or droop speed control, isochronous or
droop load control, and load sharing.

Chevron Corporation 500-37 March 1996


500 Steam Turbines Driver Manual

Woodward 509. The 509 is an electronic unit intended for both fixed-speed and
variable-speed drives. It would typically replace the PGPL or the PGG. A digital
unit, featuring two-out-of-three voting logic, it has a high degree of internal fault
tolerance as well as some self-diagnostic capability. It can be maintained on line,
accepts analog input commands, and communicates with DCS’s, LAN’s, and
various input/output devices.
The control output of the 509 is to an electric servomotor, e.g., the Woodward TM
25, which operates steam-control valves either directly or, with the appropriate
interface, through a hydraulic servosystem.
A point often overlooked in the quest for speed of response is that the governor,
even the old mechanical/hydraulic flyweight types, is usually the fastest part of the
control loop. Trying to increase speed without speeding up the servo is pointless.
However, speeding up servos can lead to instability.

542 Control Strategy


There are several control strategies but only two basic parameters to control in
steam turbine operation, speed and load. On occasion, secondary functions, e.g.,
turbine pass-out flow or pressure, also need control.

Speed Control
The most common control mode in refinery applications is speed. Speed has two
subsets, isochronous and droop.
Isochronous Control. Isochronous means constant speed or, more precisely,
constant speed irrespective of any other consideration, such as load. Once the
governor has been set for a speed, nothing changes the speed indication of the
machine in the steady state condition.
A short-term speed “excursion” may occur, however, when load changes are faster
than governor response time. NEMA SM23 classifies the allowed “excursion” for
complete instantaneous load loss from 100% speed and load.
Figure 500-33 shows the isochronous function graphically.
Droop Control. Droop control is more complex. In this mode, speed changes
(droops) slightly with load variation (increase). Droop control is used for one of
two reasons: to maintain speed stability on a driven machine with poor stability
characteristics or to gain control over load when the permitted speed band is
extremely narrow or dictated by an outside influence such as the frequency of the
local utility. See Figure 500-34.

Load Control
Load control is the normal way to run electric generators. The two basic types are
droop control and direct load control.
Droop Control. For an isolated generator, control can be either isochronous
(Figure 500-33), or droop (Figure 500-34). The isochronous condition is a special

March 1996 500-38 Chevron Corporation


Driver Manual 500 Steam Turbines

subset with zero droop, and the droop is not necessary unless the goal is to parallel
two generators, either pairing them with each other or pairing one with the utility
company.
The basic principle in parallel operation is to make one generator the base-load
machine and to make the other the load-swing machine. One machine (or the
utility) dictates the speed by being on isochronous control. This machine takes the
load swings. The other generator is on droop control and supplies a fixed load. This
load can be set at any position from 0% to 100%. This mode of control is generally
referred to as “droop isochronous.”
Droop isochronous control is illustrated in Figure 500-35. The base-load gener-
ator is set up to take, for example, 90% of its rated load. Because the speed is set by
other means—either by the utility or by the swing machine—it cannot deliver any
other load unless the speed changes. Because the governor output position is unique
to this set of conditions, which in turn puts the steam admission valve in a unique
position, the steam flow is fixed.

Fig. 500-35 Droop Control Dual Machine

If the total load is greater than 90% of the base-load machine rating, the swing
(isochronous) machine assumes the difference. However, if the total load is less
than the set point on the droop-controlled machine, the swing machine unloads
completely and the generator “motors” the turbine. This could be catastrophic, as
the turbine can overheat and be damaged unless a supply of “cooling steam” is
maintained. Fortunately, cooling steam is usually available and the generator often
has and anti-motoring protection or trip.
This system works well enough and can be performed by any NEMA C governor,
such as the PPG. With such governor types, any number of machines can be
controlled in parallel, but only one machine can be isochronous. All the others must
be on droop control but can be on a different droop percentage.

Chevron Corporation 500-39 March 1996


500 Steam Turbines Driver Manual

There are two major inconveniences in using dual-machine droop control:


• If the base-load machine separates from the swing (utility), a step change
occurs in the load the base machine sees, and a speed and frequency change
occurs on the droop-controlled machine. If the base machine picks up load, its
speed or output supply frequency decreases (droop). If the load is dropped,
speed rises. The changes may present difficulties for the electric machinery on
the dependent system or for the two generators that must be re-synchronized.
• Also, the total load swing of multiple-droop units is limited to the capacity of
the isochronous machine. Beyond that, one of the droop machines may try to
exceed full load or may reach zero load. The protection systems described
above are usually in place, which means that one or more generators will prob-
ably go off line.
Modern electronic governors give us some options in dealing with the limitations of
mechanical governors.
Direct Load Control. With the addition of load control, multiple units can run in
parallel on isochronous control. A detailed description can be found in supplier
manuals such as Woodward manual 01740B Power Management. A brief descrip-
tion follows.
The system has two critical components, a load sensor and a summing junction.
The load-sensor task is usually carried out by voltage and current transformers on
the generator output line(s). The signals are reduced to a simple signal; for
example, 0-6 volts DC would represent 0% to 100% generator output (kilowatts).
This signal is fed via a bridge circuit to a summing junction on the governor. The
summing junction also receives input from the steam-turbine speed sensor (tooth
wheel and pickup) and the speed setting input. The output of the summing junction
is used to reset the governor, and the governor output repositions the steam-admis-
sion valves through the servosystem.
Introducing the load signal to the summing junction biases the set point to make the
governor pick up load. The signal is fed through a bridge circuit to permit addi-
tional control inputs (biases). See Figure 500-36.
Two or more systems can be balanced through the bridge and each held at a fixed
percentage of its rating. If the total load changes, all units vary their actual output to
stay at the same percentage of total.
Electronic governors with these capabilities are on the market from several
suppliers such as Woodward and Tri-Sen.

March 1996 500-40 Chevron Corporation


Driver Manual 500 Steam Turbines

Fig. 500-36 Balanced Load Bridge. Courtesy of Woodward Governor Company.

543 Control Stage


The governor does the actual controlling by positioning the inlet steam control
valves.
Small turbines normally use a throttling control method. Unfortunately, throttling
reduces cycle efficiency because the available enthalpy drop is reduced. See the
Mollier Chart in Figure 500-5. A typical enthalpy drop is shown by the condition
line AB. The throttling process is typically AC, reducing the work process to CB′.
The lost energy appears as additional temperature (quality) in the exhaust steam.
The effect of throttle control on efficiency is shown in Figure 500-37. At low power
output, steam usage is disproportionately large and efficiency low. The Willens line
represents the best performance achievable, theoretically, given the blade and
nozzle design, operating speed, and supply steam properties in question. Clearly,
any losses puts the operating point above the Willens line.
An ingenious way around this problem eliminates over 90% of the loss in most
cases. The inlet flow is split into several parallel paths. See Figure 500-9. Each path
flows through a valve before reaching the nozzles. The valve opening is sequenced
so that only one valve at a time is actually throttling. The other valves are either
fully open or fully closed. No valve starts to open until the preceding valve is fully
open.
The sequence is preset and the valves are opened in turn by a rack or a camshaft.
Typically there are 3 to 7 valves. The most efficient arrangement is to have the
smallest valve modulating at normal operating conditions.

Chevron Corporation 500-41 March 1996


500 Steam Turbines Driver Manual

Figure 500-38 shows the effect on steam flow and power. Compare it to
Figure 500-37. Without throttling losses, the Willens line is straight. By increasing
the subdivision of the inlet flow, the turbine can operate more efficiently over a
wider range of conditions.

Fig. 500-37 Single Throttle Control Fig. 500-38 Throttle Control Alternate Method

544 Emergency Controls


Overview
In the event of a sudden loss of load, e.g., a coupling failure or a generator circuit
breaker trip, a turbine rapidly accelerates. NEMA SM23 lays down the limiting
speed excursions for loss of load but does not address loss of driven inertia. Even if
NEMA SM23 is fully satisfied (107% speed peak), a coupling break can cause very
rapid acceleration, and runaway speed—1.5 to 2.5 times design speed—can be
reached in 2 to 5 seconds.
Immediate intervention is required to prevent a serious self-destruct incident. In
extreme cases, the turbine casing is usually damaged in some way to allow a pres-
sure release of steam. Blades are typically torn off, rotor discs centrifugally dilate
and plastically deform, shafts bend or break, coupling hubs and shaft ends are
ejected several hundred feet. Such self-destruct events start as low as 140% design
speed for machines such as generators.

Overspeed Protection
Traditional design of protective systems was based on a spring-loaded bolt set
eccentrically in the rotating shaft. Above the set speed, centrifugal force dislodges
the bolt against the pre-loaded spring. The dislodged bolt, now protruding from the

March 1996 500-42 Chevron Corporation


Driver Manual 500 Steam Turbines

shaft, trips a latch, which in turn allows a quick-acting valve to close under spring
pressure.
Some systems are all mechanical, some partly hydraulic. In either case, the entire
chain of events takes place in about 0.1 to 0.5 seconds and is usually sufficient to
prevent damage.
More recently, electronic systems have been developed in which the eccentric bolt
and latch are replaced by a toothed wheel and magnetic pickups. Electronic
processing of the signal sends a command to the quick-closing valve which closes
under spring pressure in the traditional manner. The electronic part of the sequence
is very fast (elapsed time below 0.01 seconds), but the controlling factor is the time
the valve takes to close (0.25 to 0.5 seconds).
Sentinel Valves. If the condenser vacuum is lost, the exhaust pressure of a steam
turbine rises rapidly, exacerbated by the speed control, which increases the steam
flow to compensate for loss of efficiency and power.
The pressure rating of the exhaust hood is usually low (30 to 50 psig). A pressure-
relief valve sized for maximum flow must be present to supply overpressure protec-
tion.
For many years, manufacturers fitted a sentinel valve on the exhaust casing, rigged
to blow a whistle at approximately 15 psig. However, the sentinel valve is not a
legal protective device in any country in the world and must never be used as such,
nor must it be present alone. There are, in fact, many locations where sentinel
valves are illegal even if used in conjunction with a pressure-relief valve.

550 Operation

551 Slow Roll and Warm Up


Start-up of a steam turbine is a critical operating phase. Inadequate procedures can
easily cause reliability problems. A turbine can be started in very different ways,
depending on its size, number of stages, and inlet steam temperature.
See “Best Practice to Eliminate Turbine Slow Rolling” in the Appendix E of the
Utilities Manual.
Three potential problems arise from poor warm up or rapid starts of the turbine:
• Thermal Transients: Thermally induced stresses in heavy or thick metal compo-
nents can result in fatigue cracks after surprisingly few start-ups.
• Differential Expansion: Differing warm-up rates of components such as
casings and rotors can cause loss of critical clearances.
• Condensate: Accumulation of condensate in the inlet piping or in the casing
can cause severe impact damage if the turbine is allowed to accelerate beyond
slow roll.

Chevron Corporation 500-43 March 1996


500 Steam Turbines Driver Manual

Thermal Transients
Heat takes time to soak through any material, metals included. The property that
describes this behavior is called thermal diffusivity. This property has dimensions
ft2/hr and is defined in the following equation:

α = k/(δc)
where:
α = thermal diffusivity

k = thermal conductivity
δ = density

c = specific heat
Some typical values are:
0.5% Carbon steel α = 0.57
Copper α = 4.34
Aluminum α = 3.48
Glass α = 0.05
12% chrome steel α = 0.28 (typical)

Note the wide range of values. An in-depth study is somewhat beyond the scope of
this manual. However, note that the typical construction material for steam-turbine
rotors and casings has a low value. Also, note the very low value for chrome steel,
which is used from time to time in special applications.
The lower the value, the longer and slower the warm-up period required to avoid
the risk of crack propagation. Other factors affect the peak transient stress as well,
such as elastic modulus and thermal coefficient of expansion.
Typically, metal thickness of less than about 3 inches and steam temperatures below
about 700°F cause no problem. Thus a single stage turbine up to about 1000 bhp
can probably go on line cold, without warm up. No harm comes from thermally
induced stresses. However, the steam header must be properly warmed and conden-
sate removed to avoid re-entrainment of condensate into the steam flow and inges-
tion into the turbine.
By comparison, a multi stage turbine with a heavy drum-type rotor 12 inches or
more in diameter, running on 800 psi steam, may have a problem and could develop
cracks in the rotor core after a few dozen rapid starts. A rapid start is a function of
several factors and has no precise definition. However, slow-roll, warm-up periods
of several hours are often quoted by manufacturers. Take these recommendations
seriously. A slow roll ensures even heating and a straight rotor. It provides enough
time to make sure that the rotor is warmed to the core and that no thermal stresses
remain before centrifugal stresses are added by rotation at speed.

March 1996 500-44 Chevron Corporation


Driver Manual 500 Steam Turbines

If a problem develops, it starts as crack growth originating at the center of the rotor.
Usually, no symptoms appear until crack growth accelerates at a dangerous rate. At
this point two symptoms become observable:
• The vibration signature changes. A strong 2X or 1X signal develops, which is
not explainable by such conventional diagnoses as misalignment, balance, etc.
• Ultrasonic examination of the rotor may reveal diametral or radial separations
(cracks) close to the rotor core.
A large cracked rotor can be repaired by hollow boring (bottle boring) to remove
the damage. Because metal at the rotor center adds little strength or stiffness, it’s
not absolutely necessary. Before the repair, however, check out the rotordynamics,
as the reduced weight pushes the first critical speed up towards the operating range.
Only very large rotors can be treated this way. Typical refinery drives are usually
too small for a bottle-bore job and must be scrapped.

Differential Expansion
Single Stage Turbines. Single stage turbine rotors are short and not likely to suffer
loss of internal clearance due to differential expansion. Typically, the rotor is
located axially by a thrust bearing at inlet or the high-temperature end. Thus, expan-
sion of the rotor, which gets hot faster than the casing, tends to increase the internal
clearance between the nozzle outlet and the rotor blade leading edge. Also, because
typical single stage turbines do not have or need tip seals on the rotor blades, close
clearance is not a concern.
We can conclude that a single stage turbine can have a fast, cold startup without
suffering internal damage from loss of internal clearance. Some thermal bowing of
the rotor may occur, but the relatively generous clearances prevent rotor-stator
contact. A short period (minutes) of rough running (high vibration) may be
observed, but shouldn’t cause concern. Most single stage machines are robust
enough to withstand vibration for a few minutes at each start.
Eliminating the slow rolling of standby machinery can save steam. Of course, the
turbine must have the configuration described above, and the driven machine must
be considered as well.
Multi Stage Turbines. Multi stage turbines with their longer rotors present a very
different scenario. Internal clearances could easily be compromised during the
warm-up transient state unless the warm up is slow. Fortunately, the warm up is
usually slow for other reasons.
Also, attempting to run at speed with a rotor that is not fully stabilized thermally
could cause imbalance and result in major damage.

Condensate
Condensate can accumulate in the casing of turbines during initial warm up. It can
also accumulate in condensing turbines.

Chevron Corporation 500-45 March 1996


500 Steam Turbines Driver Manual

Most single stage turbines have self-draining cases and do not need casing drains
unless the exhaust header is higher than the casing. In this situation, the lowest part
of the exhaust system should be steam trapped.
In a multi stage condensing machine, the last two or three stages operate below the
saturation line and are wet. If water is allowed to accumulate or pass downstream
from stage to stage, severe erosion of the blades occurs, particularly at the tips.
The wet region of the casing must be provided with drain points at each stage to
allow the condensate to exit the casing at both low pressure (during warm up) and
at normal operating pressures. The drains help prevent erosive damage.
In the wet region, the stage spacing is usually set further apart to allow adequate
water separation. If supply-steam quality or superheat is significantly reduced, the
wet zone extends upstream where the casing is ill-equipped to deal with it. Severe
erosion of blades, nozzles, and even discs and shafts can occur. On occasion, entire
rotors have to be replaced.
Casings are usually warmed by allowing a small amount of steam to flow through
the inlet nozzles and vent into the atmosphere until the exit steam is dry.
The steam flow is usually controlled by hand throttling on the steam inlet isolation.
Sometimes a small bypass valve around the inlet trip and throttle valve serves this
purpose.

552 Critical Speeds


A critical speed or resonance occurs wherever a natural frequency coincides with an
exciting frequency or running speed. Normally, turbines are designed to avoid reso-
nance in the normal operating speed range. However, most turbines have to pass
through the resonance when accelerating to operating conditions. Typically, the crit-
ical speed band is about 200-500 RPM wide. The single most important task is to
keep the speed moving—never allow it to remain constant or otherwise settle
out within that band.
Many of the newer electronic governors have start-up sequencing capability and
can be programmed to avoid undesirable conditions. Older manual systems,
however, require that the operator pay strict attention.
Blade resonances are, perhaps, even more troublesome than rotor criticals. They
arise when the nozzle passing frequency coincides with the blade natural frequency.
Again, these potentially harmful resonances do not normally occur in the operating
range. But at low speeds (slow roll) there may be some interference, particularly for
long blades in large condensing turbines.
Blades can be designed to be robust enough to run on resonance at reduced speeds.
Such resonance is not unusual for high-pressure blading but is very rare for low-
pressure (long) blades.

March 1996 500-46 Chevron Corporation


Driver Manual 500 Steam Turbines

553 Steam Conditions


Refer to the utilities manual for a detailed treatment of steam and boiler-water
chemistry issues. The following is an overview from a turbine standpoint.
Steam has many parameters:
• Pressure
• Temperature
• Enthalpy
• Entropy
• Purity
• Superheat
• Quality (dryness)
The following sections summarize how these parameters affect the performance of
a turbine.

Pressure
A turbine is rated for a given inlet pressure. If the pressure is low, less steam flows,
and control valves open to maximum positions, beyond which, rated power is not
reached.

Temperature
Small changes have little effect. At constant steam mass flow, the best efficiency
speed increases proportionately with the square root of the absolute temperature,
and the best efficiency power increases in proportion to the absolute temperature.
The limiting mass flow at full throttle, however, reduces proportionately with the
square root of the (increasing) absolute temperature.

Enthalpy
Enthalpy is the total heat or energy potential of the steam.

Entropy
Entropy describes the usefulness of the steam’s enthalpy.

Purity
Purity refers to the level of contamination. Ideally, contamination is minimal, but
often the level is high enough to cause several problems:
• Chlorides come from impure feed water and are typically carried over from the
boiler as liquid spray, a fault condition, through the superheater. There they dry
out and become solid particulates in the steam stream. Finally, they get into the
turbine, where they re-dissolve at the transition zone, causing chloride stress
corrosion and pitting of the blades. The typical 12% chrome steel blades are
very susceptible. Stress corrosion cracking may also occur in high-stress areas
such as disc keyways and shrink fits.

Chevron Corporation 500-47 March 1996


500 Steam Turbines Driver Manual

• Silicates come from poor demineralization plant performance. Many modern


systems use reverse osmosis, which gives a relatively poor silicates purity and
no scope to adjust. In critical cases, some form of polishing is needed.
As the steam saturation temperature in the boiler increases, silicates in the
boiler water show an increasing tendency to vaporize. The vapor passes
through the superheater and later precipitates in the turbine, leaving a hard
scale. Although the scale is metallurgically harmless, it causes poor aerody-
namic performance and can be extremely tenacious. The only sure cure is to
water jet or glass bead blast the rotor during a teardown. Most steam turbine
manufacturers call for steam silicate contamination to be under 20 ppb for 800
psi systems. Higher pressures call for lower figures (e.g., 1500 psi systems may
need 5 ppb or lower). See the utilities manual.
• Carbonates come from boiler carryover, poor boiler solids control, and CO2
contaminated feedwater. They leave a soft deposit that can quickly choke
unused nozzle block sections. These solids are harmless, apart from their
temporary effect on efficiency and power output, and are easily removed by
water washing.

Superheat
Superheat is the margin of temperature over saturation. (Otherwise the term is
synonymous with temperature.) Ideally, superheat is sufficient to keep all but the
last few stages above the Wilson line (97% dry). Because many systems in the
company do not achieve this standard, critical components erode and must be
rebuilt unnecessarily and at great cost.

Quality
Quality refers to the dryness of the steam. Conventionally, 90% quality means 90%
dry saturated steam plus 10% by weight of water at saturation temperature.

554 Shaft Seals


The purpose of shaft seals is twofold:
• To keep steam inside the casing.
• To keep air out of the vacuum condensing system.
There are two basic seal types:
• Labyrinth
• Carbon ring

Labyrinth Seals
Labyrinth seals are probably the most common seals on large turbines. The pressure
drop is created by allowing leakage through small clearances. See Figure 500-26
for a typical arrangement.

March 1996 500-48 Chevron Corporation


Driver Manual 500 Steam Turbines

Figure 500-27 shows a typical seal support system. Each manufacturer has a
different way of arranging the parts. For details on each system, refer to the maker’s
manual.

Carbon Ring Seals


Carbon ring seals are more common on older equipment and small back-pressure
units. They function in much the same way as labyrinth seals but have segmental
carbon rings instead of steel or bronze strips. Clearances are tighter because the
carbon rings are much wider than the single steel strips. However, there are fewer
rings, and each has a higher pressure drop.
Support systems are similar to those for labyrinth seals.
Carbon ring seals have one major problem. They have a tendency, in standby equip-
ment, to wire draw or steam cut the working faces if there is any leakage. In a unit
held ready for APS, this can be a major issue because of backleakage from the
exhaust steam header.

555 Seal Support Steam Systems


The purpose of the seal steam system is to control the steam leakage and to prevent
ingress of air to condensing systems. Figure 500-27 show a typical arrangement,
although most real systems vary somewhat in the details. Always refer to the manu-
facturer’s manual.
Supply steam in Figure 500-27 comes either from the turbine casing, selected to
give an appropriate pressure, or from the steam header. The header supply is used
only during start up, when casing pressure is too low. As an alternative to seal
steam drawn from the casing, many manufacturers use leak steam from the steam
seals of the inlet control valve rack.
An eductor is sometimes used to draw off any excess flow and thus prevent escape
of seal steam close to the turbine bearings (see Figure 500-27). This action draws in
air at the seal, and then the mixture of air and steam is sent to the seal steam
condenser. Note that this is not the main condenser, which needs to run as free as
possible from air ingress (incondensibles).

560 Instrumentation
Steam Path
Instrumentation on a single stage turbine should include, at minimum, pressure and
temperature of the inlet and exhaust steam. If the pressure downstream of the inlet
control (throttle) valve is included, efficiency monitoring is straightforward, as long
as the exhaust steam is not in the condensing state.
Multi stage turbines should have temperature and pressure at inlet and at the first
stage. This point is also known as the wheel chamber and is actually the pressure
downstream of the control stage.

Chevron Corporation 500-49 March 1996


500 Steam Turbines Driver Manual

The first stage pressure is a useful indicator of blade condition. When used in
conjunction with a mass flow indication or an alternative, such as known power
output, an increase in first stage pressure indicates increase of flow (loss of blades,
fouling, etc.).
On backpressure turbines, exhaust temperature and pressure are needed for effi-
ciency monitoring.
On condensing turbines, efficiency is difficult to monitor, as steam quality indica-
tion is also required. Because no such instrumentation is available, a condensing
calorimeter is used. But it yields variable test results.
The above and other instrumentation is not in any way unusual. For details refer to
the Instrumentation and Control Manual.

570 Auxiliaries

571 Surface Condenser


See “Troubleshooting Surface Condenser Vacuums” in Appendix L of the Utilities
Manual.

572 Lube and Control Oil Systems


Refer to the General Machinery Manual.

573 Piping Systems


Refer to the Piping Manual.

580 Rerates
Rerating covers any physical work done on the turbine which either changes the
steam inlet capability or changes the output characteristics (speed, power, effi-
ciency, reliability). Each case must be assessed on its own merits, and assistance
from the manufacturer or a specialist may be necessary.
Here are a few examples of the sort of work that can be done.

581 Single Stage Turbines


Generally, single stage turbines use standard, off-the-shelf designs, with very little
that is engineered for a specific application. You may have to refer to the maker’s
catalog or contact the maker directly to establish the maximum rating of the frame
size. The maker can indicate which components keep performance below the frame-
size maximum. A competent engineer can then address those components.

March 1996 500-50 Chevron Corporation


Driver Manual 500 Steam Turbines

Nozzle Block
Typically, the nozzle block can be changed to bring power up to the frame-size
maximum. One possibility—re-drilling the nozzles to enlarge them—is not recom-
mended because it increases the bending stresses on the blades appreciably and
requires that a maximum nozzle diameter be established and the blade stress recal-
culated. A user probably wouldn’t want to risk running above the blade-stress expe-
rience of the existing design, especially without proof of reliability at those stress
levels.
A better solution is to use more nozzles—as many as the steam-chest dimensions
allow. They are probably available in a standard block from the maker. The addi-
tional nozzles increase the power and torque of the turbine without changing the
blade stress or the optimum speed of the turbine.
In extreme cases, efficiency could be affected. Also check the coupling and driven-
equipment stressing.
It will not usually be necessary to revisit rotordynamics issues.

Blades
Usually modifying blades on a single stage unit accomplishes very little. It may
help to change from a single-row to a two-row Curtis arrangement if the operating
speed is less than 60% of the turbine best-efficiency speed. However, in addition to
a rotor changeout (or at least a disk change), this arrangement requires the installa-
tion of a reversing blade row between the two moving rows.
Such a modification is fairly straightforward in most respects, unless it requires
machining of the case to locate the new blade row. In this event, a casing stress anal-
ysis is needed or, at minimum, a pressure test to 1.5 times the casing-design pres-
sure. Note that the actual test pressure selected is a function of the design
temperature and may in fact be much higher than 1.5 times actual operating pres-
sure.

Change of Steam Properties


Changes of stream properties within the frame capability are permissible and
straightforward.
Increasing pressure and superheat has a major effect. Power increases according to
P × T0.5 where P is inlet absolute pressure and T is inlet absolute temperature. Best
efficiency does not change noticeably, but the best efficiency speed increases in
proportion to T0.5.
Check that pressure and temperature design values are not exceeded, or complete a
stress analysis and/or pressure test. Increases in flow beyond the manufacturer’s
limits are frequently easy to achieve. But beware; excess flow is likely to overload
the thrust bearings and couplings.

Chevron Corporation 500-51 March 1996


500 Steam Turbines Driver Manual

582 Multi Stage Turbines


Unlike the single stage turbine, a multi stage turbine is usually an engineered
product. Many components are already operating close to their maximum capa-
bility. Do a complete and careful analysis, including all components, before
attempting any design changes.

Casings
A casing is designed as a pressure containment and, whether cast or fabricated, is
usually limited by its stiffness, not its strength.
Pressure tests are usually carried out during manufacture and not repeated there-
after. The “new” pressure test is calculated to give a stress equivalent to 1.5 times
design at the design temperature. It is often close to twice actual operating pressure.
A casting often has a potential pressure capability much higher than the application
calls for. Take care in a pressure rerate situation. Any recalculation shows only
gross excess capability. A new higher test pressure representative of the new oper-
ating pressure could reveal a casting fault not uncovered during original testing at a
lower pressure. Alternatively, a careful NDT such as U/T or x-ray could reveal
hidden faults. Re-verification by hydrotest is recommended.
The hydrotest will probably have to be staged with internal baffles to test the HP
end adequately without over-pressuring the LP end.

Nozzles
Multi stage turbine nozzles are treated much like those in single stage turbines.
Extra nozzles can often be added by substituting nozzle blocks in the steam chest.
The control-stage blading probably does not need changing, but the downstream
(Rateau) staging has a bending stress increase. The Goodman diagrams must be
redrawn and evaluated for those stages. See the subsection on blades. Also, the
admission valves probably need to be enlarged. Velocity through the valves should
not exceed about 120 feet per second.

Blades
The control stage is not usually a limitation because it is designed to run partial
admission. Of the remaining stages, the first and last rows frequently pose a
problem for different reasons.
First Pressure Stage. The first pressure stage is designed to run at or close to full
admission. Consequently it has very little spare open area and may be the first stage
to choke (sonic velocity in the throat). If further pressure is allowed to build up
ahead of this stage to “force” more steam through the turbine, the function of the
control stage is affected; power is lost and control may become marginal.
In many cases, choking on the first pressure stage is a turbine casing limitation. If
so, an uprate can still be achieved by removing this row, making the second pres-
sure stage the first. This new first pressure stage now has about two times higher
mass flow capacity than it was originally designed for. The remaining stages
distribute the pressure drop somewhat evenly, but tend to load up the last stage.

March 1996 500-52 Chevron Corporation


Driver Manual 500 Steam Turbines

Most likely all the pressure staging must be redesigned to take full advantage of the
higher pressures. This is a difficult task. As the enthalpy drop per stage becomes
greater, one or more stages may approach choke, and the turbine may become unre-
liable and inflexible.
Last Pressure Stage. The last stage has the longest blades and the highest proba-
bility of limiting blade stress. Increased steam flow increases the bending (vibra-
tory) stress. It is essential to revisit the Goodman diagrams for all stages, especially
the last row or two. Having Goodman factors (ratio of permissible alternating stress
to actual alternating stress) above about ten is best, although five is serviceable.
The last stage, however, can present special problems. If the projected pressure
drop across the last stage increases, that increase operates largely on the rotor
blades, increasing the reaction and developing additional thrust on the rotor and its
thrust bearing. There may also be a random distribution of energy in the form of
buffeting, which can excite blade vibrations even when the turbine is running well
away from resonance conditions. These loads are very difficult to estimate.
Typically, exit axial velocity component should not exceed about 350 feet per
second on condensing turbines and 250 feet per second on backpressure turbines.
The last stage (or two) are probably limiting on condensing turbines. Backpressure
units, on the other hand, may have some margin.
If the last stages must be re-bladed, involve the OEM and select blades that have
been adequately proven by tests or field experience. Larger blades lead to higher
root loads and often necessitate replacing the rotor disk as well.

Couplings
Coupling type changes are currently fashionable for several valid reasons.
Diaphragms couplings, usually the multiple-membrane type, also known (incor-
rectly) as shim pack, are readily available, competitively priced, and well manufac-
tured. The overall dimensions are comparable to the gear-tooth types they typically
replace, making for easy conversion.
Peripheral velocities are comparable between gear tooth types and multi-membrane
types. However, windage can cause overheating if no oil spray is used. Pay atten-
tion to anti-windage baffles and ventilation if the surface velocity exceeds about
300 feet per second.
Weights are comparable. Therefore lateral rotordynamics issues are not a problem.
Some types, notably the Zurn Ameriflex, are heavy and may depress the second crit-
ical speed slightly. Make sure not to allow critical speeds to encroach on the oper-
ating separation margin.
Torsional stiffnesses are considerably different; torsional analysis may need re-veri-
fication.

Chevron Corporation 500-53 March 1996


500 Steam Turbines Driver Manual

Change of Steam Properties


Increased steam pressure and temperature operate as they do for single stage
turbines. However, the design margin over rated conditions is not usually enough to
make it an attractive proposition.

583 Rotordynamics
Any changes to the speed range or to rotating masses or stiffnesses calls for another
look at rotordynamics. This is a job for the specialist. The OEM is a good choice.
Consultants are readily available, but costs are considerable. CRTC has some capa-
bility. Contact the Machinery Group.
Here are some indications of the directional result of certain changes:
• Added mass to rotors reduces first and second critical speeds.
• Added coupling mass depresses second criticals but have little effect on first
critical.
• Increased bearing stiffness, support stiffness, oil viscosity, and dynamic
damping all increase critical speeds.
• Coupling-type changes impact the torsional frequencies that are critical for
drives with electric motors, generators, gears, and reciprocating equipment.
• Uprates on turbines with lower-half steam chests can cause rotordynamic insta-
bilities.
For a more complete description see the Rotordynamics section of the Compressor
Manual.

March 1996 500-54 Chevron Corporation

You might also like