You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www

.researchgate.net/publication/267503886
Rotordynamics of a Two-Phase Flow Twin Screw
Pump
Conference Paper in Journal of Engineering for Gas Turbines and Power June 2012
DOI: 10.1115/GT2012-69965
CITATION
1
READS
257
2 authors:
Ameen Muhammed
Baker Hughes Incorporated
5 PUBLICATIONS 24 CITATIONS
SEE PROFILE
Dara Childs
Texas A&M University
188 PUBLICATIONS 2,107 CITATIONS
SEE PROFILE
All content following this page was uploaded by Ameen Muhammed on 03 February 20
16.
The user has requested enhancement of the downloaded file. All in-text reference
s underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them
immediately.
Thank you for using the
University at Albany s
Interlibrary Loan Service
NOTICE WARNING CONCERNING COPYRIGHT RESTRICTIONS
The copyright law of the United States (Title 17, United States Code) governs th
e
making of photocopies or other reproductions of copyrighted material. Under cert
ain
conditions specified in the law, libraries and archives are authorized to furnis
h a
photocopy or other reproduction. One of these specific conditions is that the ph
otocopy
or reproduction is not to be "used for any purpose other than private study, sch
olarship,
or research." If a user makes a request for, or later uses, a photocopy or repro
duction for
purposes in excess of "fair use," that user may be liable for copyright infringe
ment. This
institution reserves the right to refuse a copying order if, in its judgment, fu
lfillment of the
order would involve violation of copyright law.
Questions?
Call 442-3613 from 10:00 ~ 4:30 (weekdays)
or
Send email to libill@albany.edu
Ameen R. A. Muhammed
Graduate Research Assistant
e-mail: ameen@turbo-lab.tamu.edu
Dara W. Childs
Leland T. Jordan Professor of
Mechanical Engineering
e-mail: dchilds@tamu.edu
Turbomachinery Laboratory,
Texas A&M University,
College Station, TX 77843-3123

Rotordynamics of a Two-Phase
Flow Twin Screw Pump
Twin screw pumps are positive displacement machines. Two meshing screws connecte
d
by timing gears push the fluid trapped in the screw cavities axially from suctio
n to discharge.
Available steady state hydraulic models predict pump performance and axial
pressure distribution in the chambers in single- and two-phase flow conditions.
However,
no model is available for their rotordynamics behavior. Due to the helix angle o
f the
screw, the pressure distribution around the rotor is not balanced, giving rise t
o both
static and dynamic lateral forces. The work presented here introduces a starting
point for
rotordynamic analysis of twin screw pumps. First, we show that the screw rotor s g
eometry
can be represented by axisymmetric beam elements. Second, we extend the steady
state hydraulic model to predict both the static and dynamic lateral forces resu
lting from
the unbalanced pressure field. Finally, hydraulic forces are applied to the roto
r to predict
static, synchronous, and nonsynchronous responses. Predictions of the dynamic pr
essure
were compared to measurements from the literature and were found to be in good a
greement.
[DOI: 10.1115/1.4023490]
Introduction
Screw pumps have been around for a long time. The earliest
version is the famous Archimedes wheel still in use for raising
water. Historically, the domain of twin screw pump applications
was highly viscous liquids. Since the early 1990 s, pump users
realized they can be used successfully for multiphase pumping
and less viscous liquids [1]. This was particularly attractive for
the oil and gas industry, where both viscosity and gas volume
fraction (GVF) of an oil field can vary widely over its life. Moreover,
this extension of the screw pumps domain of application
coincided with an increased interest in off-shore production where
the versatility of multiphase pumps is especially evident. Pumps
that accommodate high GVF can be placed on the sea bed near
the well head to raise the pressure of the mixture of crude oil,
water, and gas for direct transport to a central platform for processing.
Shippen and Scott [2] showed the benefit of this arrangement
as compared to having a satellite platform to handle the two
phases separately before delivery to the central platform.
Naturally, interest in these new multiphase capabilities drove
more research in the twin screw pump. Vetter and his group in the
University of Erlangen conducted pioneering research in screw
pumps in the 1990 s. Vetter and Wincek [3] developed a steady
state model for two phase flow. They simplified the action of the
pump by modeling it as a series of disks moving continuously
from suction to discharge. The disks represent the pump s chambers.
Their research objective was to define pump performance for
single-phase and two-phase operations to serve in the pump
design process. Their model s basic assumptions were:
(1) fully developed steady state flow.
(2) all clearances connecting the chambers are filled with liquid
phase only due to the centrifugal effects of the screw rotation.
Their experiments confirmed the validity of the
assumption for a GVF up to 80%.

(3) ideal gas and isothermal compression, due to the higher


heat capacity of the liquid phase.
Their model performance predictions agreed with their tests
adequately. They reported measured and calculated screw rotor
static deflection resulting from unbalanced hydraulic pressure and
included the effect of clearance eccentricity due to rotor deflection
on the pressure and flow calculations. Their tests also concluded
that 80% of the leakage goes through the circumferential clearance,
while only 20% goes through the more geometrically complex
radial and flank clearances. Prang and Cooper [4] later used
this empirical observation to simplify their steady state model.
They treated the circumferential clearance as a short seal and used
a Moody friction coefficient. Their model was simpler than Vetter s
but agreed satisfactorily with their test results. Vetter et al.
published additional test results in Ref. [5] and compared them
against results of the model in Ref. [3]. The test data included
dynamic pressure measurements of sensors distributed axially
along the screw length. These measurements will be used to validate
the model presented in this study.
Rausch et al. [6] employed a different modeling approach.
Instead of the disk model adopted by Vetter et al. [5] and Prang
and Cooper [4], they treated the chambers as an open thermodynamic
system. They used mass and energy conservation equations,
along with initial and boundary conditions of suction,
discharge, and GVF, to build a differential equation model that
can be solved by time-integration methods until steady state is
achieved. They relaxed the isothermal compression assumption
for an adiabatic process since their model included energy balance,
but they neglected wall friction and kinetic energy. The
main advantage of their model over the previous ones is that it
does not assume a steady-state operation, thus it can be used to
model transient performance such as in slug flow. However, their
comparison with testing was poor especially in low speeds and
high GVF.
Martin created a simple model that can be used to predict performance
for operational purposes, without the need of special
information that pump designers might be unwilling to share, such
as clearances [7]. Xu and Chan also examined performance at
very high GVF from an operational point of view [8,9]. Rabiger
et al. [10] and Rabiger [11] gave more rigorous modeling of screw
pumps at high GVF. His model is a more complete version of the
model from Ref. [6]. As with Rausch et al. [6], Rabiger s model is
based on describing the chamber as an open thermodynamic system
moving from suction to discharge expressed by ordinary differential
equations for mass and energy conservation. However,
he gives a more elaborate clearance leakage model, including the
effects of friction and screw rotation. A key addition to his model
is the relaxation of the liquid-filled clearance assumption. He
assumes a homogeneous gas/liquid mixture in the leakage path
and uses linear interpolation of the two phases to determine the
density and viscosity of the mixture. He also coupled his model
Contributed by the International Gas Turbine Institute (IGTI) of ASME for
publication in the JOURNAL OF ENGINEERING FOR GAS TURBINES AND POWER. Manuscript
received July 14, 2012; final manuscript received November 24, 2012; published
online May 20, 2013. Editor: David Wisler.
Journal of Engineering for Gas Turbines and Power JUNE 2013, Vol. 135 / 062502-1
Copyright VC 2013 by ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
with a thermal model of the rotor. His theoretical and test results

were in good agreement.


Twin Screw Pump and Rotordynamics. The literature is
almost silent when it comes to the rotordynamics aspect of screw
machines, both pumps and compressors. Traditionally, there was
no practical need to study their rotordynamics because screw
pumps are always running below the first critical speed. This circumstance
might change as the pump domain of applications
extends to cover less viscous fluids allowing for faster operation.
Moreover, some users and researchers voice concerns about field
vibration problems [9]. A frequent complaint is mechanical-seal
failure [1,2], which could be vibration related. Finally, considering
the huge capital cost invested in subsea installations, an
adequate rotordynamic model is needed to understand and predict
the vibration of these machines under working conditions, especially
since screw pumps mostly run on roller bearings, which
provide little damping. As for the damping provided by working
fluid, it has been neither studied nor confirmed by test results.
The fact that the hydraulic forces around screw pumps are not
balanced has been known since the 1960 s. Pump designers have
approximate formulas [12] to estimate those forces and insure that
rotors are rigid enough not to touch each other or the liner. Most
of the hydraulic models mentioned above take into account the
change in the leakage due to the static rotor deflection resulting
from the unbalanced hydraulic forces. However, no attempt has
been made to model or measure these dynamic forces and understand
their interaction with the rotors. Strong evidence shows that
the hydraulic forces have oscillatory components. First, the
dynamic pressure measurements reported by Vetter et al. [5] and
by Rabiger et al. [10] both show harmonic variation. Second,
screw compressors bearings are known to be subjected to
dynamic forces from the compression process (Ref. [13], p. 63).
Although they are different machines, screw compressors and
screw pumps are both positive displacement machines. In fact,
Stosic et al. argues in Ref. [14] that the basic modeling techniques
can be extended from screw compressors to screw pumps without
significant modifications. While, this might not be entirely accurate,
similar trends are expected.
The objectives of this research are: (1) Develop a structural
dynamic model of the screw rotor, (2) develop a hydraulic model
for the transient pressure field around the screw as a function of
its rotational angle, and (3) study the response of the structural
model to the transient hydraulic forces.
Structural Model of Screw Rotor
In this section, we will first discuss lateral rotordynamics, then
torsional, and finally lateral-torsional coupling due to the timing
gear.
Figure 1 is a simplified solid model of a typical single-thread
screw rotor in a twin screw pump. The rotor is for a 75 kW pump
running at 1750 rpm.
The screw section of the solid model and its cross sections are
shown in Fig. 2. Ls is the screw section length. Dout and Din are
outer and inner diameters of the screw thread, respectively. h is
the screw thread pitch. At any given axial location, the cross section
of a single-thread screw is composed of the two half circles
shown in Fig. 2. The orientation of the two half circles cross section
goes through a complete revolution in a pitch length.
Table 1 is a list of the first five free-free modes of the general finite
element (FE) screw rotor solid model in Fig. 1 in two orthogonal
directions. The analysis is performed in ANSYS. The X and the
Y free modes are almost identical showing that the rotor is nominally

isotropic and can be modeled by an equivalent simpler axisymmetric


structure.
A schematic of the axisymmetric model is shown in Fig. 3. The
analysis was performed using a circular Timoshenko beam finite
element model. The geometry of the thread is represented by two
circular cross section layers of different diameters. One layer of a
95mm diameter for the mass properties and the other is a 69mm
diameter for the stiffness properties.
The equivalent mass diameter is simply the circular diameter
that results in the same volume of the thread section. Given the
shape of the cross section shown in Fig. 2, an equivalent-stiffness
circular cross section is expected to have a diameter only slightly
larger than the inner diameter. The exact value is found by trial
and error to match the free-free modes of the nonaxisymmetric
model. Table 2 compares the free-free modes of the axisymmetric
and nonaxisymmetric models. The small errors show that an
equivalent axisymmetric model can be used to model the nonaxisymmetric
thread.
While the nonaxisymmetric cross section was found to be
adequately modeled by a constant circular cross section, the
thread is not necessarily dynamically balanced. To investigate the
effect of the nonsymmetric cross section on the dynamic imbalance
response, the pitch length (h) is sliced into Nsl pieces. The
center of mass of each slice is displaced from the center of geometry
of the rotor by the length of the vector usl given by
Fig. 1 Solid model of a twin-screw pump rotor (dimensions in mm)
Fig. 2 Screw sections (dimensions in mm)
Table 1 Screw rotor free-free nonaxisymmetric modes
Free-free frequency (Hz)
Mode x direction y direction
1st 281.24 281.63
2nd 682.21 682.71
3rd 1274.9 1277.4
4th 2030.2 2032.2
5th 2913.1 2914.7
062502-2 / Vol. 135, JUNE 2013 Transactions of the ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
jusl j
4 R3
outR3
in
 
3p R2
outR2
in
 ; ffusl k
2p
Nsl
k0;1;2 ;Nsl1 (1)
Rout and Rin are the outer and inner screw radii, respectively.
The amount of the unbalanced mass for each slice (msl) is given
by
msl qVscrew
h
NslLs
(2)
q is the density of the screw section material. Vscrew is the volume
of the screw section. h is the screw pitch. Ls is the axial length of
the screw section. Nsl is the number of slices.

Figure 4 shows the predicted response due to the imbalance inherent


in the geometry of the threads. No additional imbalances
were applied. The bearings are isotropic and identical with a
6.1107 N/m stiffness and 525 N-s/m damping. The cross section
was sliced into eight pieces (Nsl8). The large amplification factor
(30.82) is expected due to the lack of damping at the bearings.
The response at the critical speed is 25mm peak to peak, suggesting
the machine cannot pass the critical speed without appropriate
dynamic balancing. However, the response at the running speed is
0.045mm at the middle of the rotor and 0.028mm at the seal at
the suction sides, which explains why the dynamic imbalance of a
single-thread twin screw pump might not be of a concern in the
current application margins.
Similar to the lateral direction, the screw thread can be treated
as an axisymmetric structure in the torsional direction. For the
same mass and stiffness screw diameters used for the lateral direction
analysis, the first three torsional modes were in good agreement
with the nonaxisymmetric torsional modes from the solid
model general FE analysis as shown in Table 3.
The high values of the torsional modes in Table 3 can be misleading.
The analysis is performed only for one rotor for matching
purposes. In reality, the torsional system consists of a motor, coupling,
and two rotors connected by timing gears.
The coupling s torsional stiffness varies depending on the
applications and design. For a soft coupling, and including the
gear tooth stiffness, the system s first torsional mode can drop significantly
as shown in Fig. 5.
Gears are a known source of torsional-lateral coupling. In most
machines, it is not an issue; however, for long coupled rotors, the
torsional lateral coupled critical speed might be appreciably lower
than the uncoupled. Rao et al. [15] and Lee et al. [16] both
Fig. 3 Axisymmetric screw rotor model (dimensions in mm)
Table 2 Axisymmetric versus nonaxisymmetric free-free lateral
modes
Mode
General FE
ANSYS (Hz)
Axisymmetric
XLTRC2 (Hz) Error
1st 277.99 277.3 0.2%
2nd 673.91 670.2 0.6%
3rd 1267.3 1280.2 1%
4th 2013.3 2030.1 0.8%
5th 2886.9 2943.3 2%
Fig. 4 Response due to cross section imbalance
Table 3 Axisymmetric versus nonaxisymmetric free-free torsional
modes
Mode
General FE
ANSYS (Hz)
Axisymmetric
XLTRC2 (Hz) Error
1st 1265 1217.9 3.72%
2nd 2344.9 2342 0.12%
3th 3153 2937.1 6.85%
Journal of Engineering for Gas Turbines and Power JUNE 2013, Vol. 135 / 062502-3
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
develop a straightforward finite element representation for two
axisymmetric rotors connected by a flexible gear tooth. The system

in Fig. 6 consists of a motor, a coupling, two rotors, and a torsional


lateral coupled gear connection. The motor is modeled as
an added mass (180 kg) and an added inertia (9 and 18 kg-m2 for
polar and transverse inertia, respectively). The coupling is represented
by an added stiffness (3106 N/m in the lateral direction
and 1104 N/rad in the torsional direction). The coupling added
mass is 26.6 kg. Both gears have a 43mm radius and a 5.4 kg
mass. Gear tooth stiffness is taken 1106 N/m.
A comparison of the lateral, torsional, and coupled lateraltorsional
system s natural frequencies is shown in Table 4.
The first lateral mode shapes for the two shafts are shown in
Fig. 7. The modes are characteristic of rigid bearings systems with
most of the vibration energy going in the shafts flexibility.
The first torsional natural frequency at 2271 rpm is the mode
affected by the coupling stiffness, where the coupling and motor
masses vibrate in a direction opposite to the rest of the system.
The torsional lateral coupling adds a new low frequency mode
at 3227 rpm. The corresponding mode shape is shown in Fig. 8,
where the interaction between the vibrations in the torsional and
lateral directions through the timing gear is evident.
Hydraulic Model of Twin Screw Pump
Having touched on some structural characteristics of screw
pump rotors, the next step is to characterize the dynamic pressure
field in a full revolution to predict the resultant forces and
response. The available hydraulic models in literature predict the
steady state pressure distribution in the chambers and the pump
performance [3,4]. In this section, the steady state models are
extended to predict the pressure field around the screw as a function
of the rotor s angular position. The dynamic pressure characterization
model is composed of the following three components:
(1) the steady state hydraulic model to predict the axial pressure
distribution in the chambers and across the screw lands, (2) the
model of the pressure buildup in the last chamber as it opens to
the discharge, and (3) the extension of the axial pressure distribution
to a pressure field around the screw and pressure field variation
with respect to the rotational angle of the rotor. Although the
outcome of the analysis is a dynamic pressure of a transient nature
(pressure field versus the rotor s rotational angle or time) the
model output is strictly steady state with results in the frequency
domain.
Steady State Hydraulic Model. The steady-state hydraulic
model predicts the pressure in the chambers and across the screw
lands given the discharge and suction pressures, the GVF at the
suction, the pump geometry, and running speed. The present
model is based on both Vetter and Wincek [3] and Prang and
Cooper [4]. For brevity, we will only list the assumptions, go over
the main equations, and provide the model s algorithm.
The main assumptions of the steady state model are:
(a) The backflow in the clearances is only liquid phase.
(b) The backflow will only be evaluated for the circumferential
clearances. The slip through the flank and radial
clearances will be assumed as an empirical percentage
of the circumferential backflow. A schematic of the
clearances is shown in Fig. 9.
(c) The leakage flow across screw lands through the circumferential
clearance will be treated as an annular seal.
(d) The gas phase behaves ideally, and the compression is
isothermal.
(e) The pressure in the chamber is constant.
Assumptions (a) and (b) are based on the experiments by Vetter

and Wincek [3]. They showed that, for a GVF below 85%, the
centrifugal effect of the screw rotation will cause the clearance to
be only filled with the higher density liquid phase. For assumption
(b), they ran nonrotating tests where they blocked different types
of clearances successively with synthetic resin and forced liquid
through the unblocked clearances by external pressure. They concluded
that 80% of the backflow occurs through circumferential
clearance. The annular seal assumption (c) is justified in light of
the geometry of screw lands. The clearance between the screw tip
and the liner is kept tight to avoid metal-to-metal contact (typical
clearance to radius ratio range is from 0.002 to 0.003). The length
to diameter ratio is rarely over 0.2. A finite-length liquid seal
model was developed for this work based on Ref. [17]
Fig. 5 Variation of first torsional natural frequency with torsional
coupling stiffness
Fig. 6 Torsional-lateral gear coupled system
Table 4 Lateral, torsional, and coupled system s natural
frequency
Lateral
(rpm)
Torsional
(rpm)
Torsional/Lateral
KGR1106 (N/m)
5636 (BW)a 2271 1721 (BW)
5707 (FW)b 17,000 3227 (FW)
5996 (BW)
5636 (BW)
6002 (FW) 5707 (FW)
5996 (BW)
6002 (FW)
aBW designates backward critical speeds.
bFW designates forward critical speeds.
062502-4 / Vol. 135, JUNE 2013 Transactions of the ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
(pp. 229 238), but the results were identical to the calculations
based on the short seal assumption. Inherent in the short seal
assumption is neglecting the axial movement of the screw tip opposite
to the backflow due to its helix shape as well as the slanted
angle of the seal along the circumference. Ideal gas assumption
(d) is a typical starting point for all models involving gas compression.
It can be relaxed for real gas properties without significant
changes if a more accurate model is desired for oil and gas.
The isothermal compression assumption is based on the large difference
in the heat capacities of the gas and liquid phases as mentioned
earlier. The experiments by Signh [18] found the assumption
valid for GVF up to 94%. Finally, the constant chamber pressure is
effectively an assumption of balanced flow. Pump users mention
that slug and unbalanced flow exist in operation especially at high
GVF and can cause vibration trouble especially in the axial direction;
however, this issue will not be addressed here.
Screw pumps are positive displacement machines. Similar to
piston pumps, they work by forcing the fluid against the back
pressure. Isolated pressure zones (chambers) are formed and outlined
by the pump liner and the meshing with the other screw as
shown in the cross sectioned pump in Fig. 10. Once a chamber is
sealed from the suction, the only communication between the
chamber and the rest of the system is through the clearances. The
slip flow is the total leakage from one chamber to the chamber
upstream through all the clearances connecting the two chambers.

For multiphase operation, the entrapped fluid in a chamber is


pushed forward by the rotation of the helical screw toward the discharge
side. Meanwhile, the inflow from the chamber downstream
exceeds the outflow to the chamber upstream, causing the pressure
to build up. The pressure buildup results in gas phase compression
and liquid phase accumulation. In a hypothetical pump running
with zero clearances, the pressure in a chamber would remain at
the suction level until it suddenly sees the discharge side and
the fluid is pushed out by the action of the screw. Neglecting the
thermodynamics, a screw pump model reduces to a mass flow balance
between the backflow through the clearances and gas compression
in each chamber. For a single-phase flow, the pressure
buildup increases linearly from suction to discharge.
Chambers are numbered from 1 to nch, where 1 is the chamber
closest to suction and nch is the chamber closest to the discharge.
Fig. 7 First two forward lateral mode shapes without torsional-lateral coupling
Fig. 8 Torsional-lateral mode shape
Fig. 9 Screw pump clearances [11]
Journal of Engineering for Gas Turbines and Power JUNE 2013, Vol. 135 / 062502-5
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
The axial velocity of the leakage flow across the circumferential
clearance depends on the pressure difference between the two adjacent
chambers, the friction across the circumferential clearance,
and the inlet pressure loss due to the sudden acceleration of the
flow entering the clearance as given by [4]
Vcclkg;i1!i
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiff
iffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Pi1  Pi
ql
2
1 n ql
B
dcc
fcc
vuuut
(3)
Pi1 and Pi are the pressures in chambers i1 and i. ql is the density
of the liquid phase since the back flow is assumed all liquid. B
is the length of the screw land (Bh/2N where h is the screw
thread pitch and N is the number of threads). dcc is the nominal
circumferential clearance height. n is the entrance loss coefficient.
fcc is the Fanning friction factor for the flow across the circumferential
clearance.
Unlike Ref. [4], where a Moody-based friction factor definition
was used, a Blasius friction model will be adopted here. The value
of the friction factor for turbulent flow is dependent on the Reynolds
number as given by
fcc nccRmcc
e (4)
ncc and mcc are constants based on the surface roughness. Re is the
axial Reynolds number across the circumferential clearance given
by
Re
2Vcclkg;i1!idcc

(5)
 is the kinematic viscosity of the liquid. Since the friction is a
function of Reynolds number, the procedure is iterative.

The total slip flow rate from chamber i1 to the upstream


chamber i is given by
Qslp;i1!i
1
ft
 Vcclkg;i1!i  Acc (6)
ft is the empirical fraction of the axial leakage flow through the
circumferential clearance to the total slip between the two chambers
through the flank, radial, and circumferential clearances.
According to Ref. [3], the estimated value of ft from the experiment
is 0.8.
From the ideal gas assumption, the gas volumes in two chambers
are related to the pressure by
Vg;i Vg;i1 
Pi1
Pi
(7)
Pi and Pi1 are the pressures in chambers i and (i1). Vg,i and
Vg,i1 are the volumes occupied by the gas in chambers i and
(i1). The volume occupied by the gas and the volume occupied
by the liquid add up to the chamber geometrical volume. For the
chamber created at the suction side, the liquid volume is given by
the GVF at suction.
VL;s 1  GVFs  VK (8)
VK is the chamber volume determined by the geometry of the
thread. VL,s is the liquid volume in the chamber closest to suction.
The chamber life time Dtlife is defined as the time period in
which the fluid in the chamber is transported axially a distance
equal to the pitch length from suction to discharge. It is given by
Dtlife
2p
Nx
(9)
X is the running speed, N is the number of threads.
The volume of the slip flow from one chamber to the chamber
upstream over the period of chamber life time Dtlife is
VL;i1!i Dtlife  Qslp;i1!i (10)
The algorithm for steady state pressure calculations is shown in
Fig. 11. Initially, the pressure in all chambers is assumed to be suction
pressure, and the volume of the liquid in all chambers is
assumedVL;s from Eq. (8). Starting from the discharge side, the liquid
volume in each chamber is updated according to Eq. (10). The
corresponding gas volume in the chambers is calculated based on the
chamber geometrical volume and the liquid volume in the chamber.
The chamber pressure is updated from the ideal gas Eq. (7). The final
chamber pressure distribution is used as the initial guess for the next
iteration. The process concludes when the maximum difference in
chamber pressures between two iterations falls below a desired tolerance.
Since the screw land is modeled as a short annular seal, the
pressure across the screw land is assumed to drop linearly.
Chamber Opening to Discharge. The process of opening the
last chamber to discharge is a complicated hydraulic process. An
approximation adequate for the purpose of this research is given
here. To understand the complexity of the process, note that it is
fundamentally a transient phenomenon. Our model on the other
hand is not a transient model. While our final goal is to find the
pressure field variation as a function of shaft rotation, we are not
interested in time dependent events; we are rather interested in the
variation of the steady state pressure due to the rotation of the
unsymmetrical screw rotor.

Figure 12 is a schematic of the conceptual stages of the process.


Before the last chamber opens to the discharge, its pressure level
is less than the discharge pressure. Once a gap starts to open
between the screw and the liner, fluid flows from the higher pressure
level on the discharge into the chamber through the tight gap.
This inflow (assumed to be all liquid) causes the chamber pressure
to rise. Meanwhile, liquid continues to leak from the last chamber
to the chamber upstream through clearances. As the process continues,
the gap area increases while the chamber volume decreases
due to shaft rotation. Finally, the pressure in the chamber reaches
the same level of the discharge, and fluid is squeezed out by the
decrease of the chamber volume.
The shaded area in Fig. 13 is the gap that opens the last chamber
to the discharge. The gap is outlined by the pump liner, the
rotors inner tip, the meshing from the other screw, and the screw
thread terminal edge. The angle between the meshing of the two
screws and the screw thread terminal edge is /. / defines the area
of the gap, and it ranges from 0 deg to 180 deg for a singlethreaded
screw.
Fig. 10 Twin-screw pump cut away section [9]
062502-6 / Vol. 135, JUNE 2013 Transactions of the ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
Manring [19] develops a model for the similar situation of piston
opening in the axial-piston swash-plate hydrostatic pump. The
following treatment is based on his work. The inflow from the discharge
to the last chamber through the gap is modeled by the Bernoulli
orifice equation
Q/gap Cd/A/gap
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiff
iffiffiffiffi
2
ql
Pd  P/n


s
(11)
Cd(/) is the instantaneous coefficient of discharge associated with
the gap orifice opening to the discharge. Its value is estimated
from a CFD model of the gap. It was found to range from 0.6 to
0.7 depending on the rotational angle. A(t)gap is the instantaneous
cross section area of the gap orifice, and P(t)n is the instantaneous
pressure in the chamber opening to discharge.
Manring [19] defines the rate of the pressure buildup in the last
chamber due to the backflow from the discharge by the following
capacitance model.
dP/d
d/

30
pX
j 0
Vnch /
Q/orifice;dQ/ slp;d!nch
dVd/
d/
 
(12)
Q/ slp;d!nch is slip flow from the discharge to the last chamber

as calculated from the slip model above. dVk=dt is the rate of


chamber reduction in volume. The effective modulus of elasticity
b 0
accounts for the presence of gas and is adopted from Ref. [20].
X is the rotor running speed.
At each angular shaft increment, the discharge chamber pressure
from Eq. (12) is used to set the boundary condition for the
steady state hydraulic model to calculate the pressure in the rest of
the chambers and over the screw lands.
The dynamic pressure measurements reported by Vetter et al.
[5] showed a spike of the pressure in the last chamber over the discharge
pressure (their test data is reproduced in Fig. 16). Such a
spike is best understood in the framework of transient hydraulics
of multiphase flow. Tullis explains in Ref. [21] (p. 250) that an air
pocket passing through an orifice causes a transient pressure rise.
The difference in densities between the gas pocket and liquid
Fig. 11 Steady state hydraulic model algorithm
Fig. 12 Schematic of the process of opening the last chamber
to discharge
Fig. 13 Gap opening to discharge
Journal of Engineering for Gas Turbines and Power JUNE 2013, Vol. 135 / 062502-7
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
phase results in an acceleration of the flow as the gas pocket
passes the orifice and a sudden deceleration as the liquid passes.
The sudden reduction in velocity causes a pressure head. Martino
et al. [22] ran a number of tests and compared their results favorably
with a simplified model that predicts the maximum head rise
DHmax in meters. The details of the pressure spike model will be
skipped for brevity.
Pressure Field as a Function of Angular Displacement. So
far, the model treats the screw pump in one dimension. Each
chamber is represented by a single isolated one-dimensional constant
pressure zone. To calculate a two-dimensional pressure field
around the rotor, the helical screw geometry has to be taken into
consideration.
Figure 14 is an unwrapped screw section. The horizontal axis
denoted by coordinate (z) is along the axial direction of the screw.
The vertical axis denoted by coordinate h is along the polar circumferential
direction (zero h is taken from the positive y axis
shown in the schematic to the right, positive h is counterclockwise).
The screw lands are the gray strips, and the chambers are
the white strips. The vertical black lines are provided to help visualize
the unwrapped geometry. They show that the screw lands
and chambers at the top are connected to the ones at the bottom
without discontinuity in the geometry. In a circular rotor, angles 0
(at the bottom) and 2p (at the top) are the same angle.
The unwrapped geometry is discretized to small squares in a
two dimensional array. This discretization should not be confused
with a finite element or finite difference solution scheme. The
model is essentially a steady state mass balance. The discretization
is for book keeping to record the pressure data around the
screw versus the angular rotation of the rotor.
With this geometry in mind, the arrows illustrate that, for a single
screw, the discharge and suction sides are hydraulically connected.
The high pressure fluid in the discharge side will flow
back following the arrows without obstruction to the low pressure
in the suction side. By itself, a single screw will not be divided
into chambers. It is the meshing of the other screw that provides
the obstruction to the flow, which creates multiple isolated constant

pressure zones. The meshing occurs on a tangent line with


clearance to inner radius ratio in the order of 0.0015. On the two
dimensional schematic in Fig. 14 it is shown as a dotted horizontal
line.
For a stationary observer, the meshing line is fixed at one location,
and the rotor is spinning. For an observer spinning with the
rotor, the situation is reversed, the rotor is stationary, and the
meshing line is spinning around the rotor 360 deg. Therefore, the
action of rotating the screw rotor is represented on the two dimensional
unwrapped geometry by advancing the meshing line on the
vertical axis.
This geometrical representation ties together the elements of
the model and provides a procedure to express the pressure field
around the rotor as a function of the angular position. Figure 15
shows the variation of the pressure in two dimensions at different
angular position of the rotor. The calculations are performed for
the pump in Vetter et al. [5]. The pump is a double-thread twin
screw pump, operating at a 15 bars pressure head, a 75% GVF at
suction, and running at 3000 rpm. The axial and circumferential
directions are along the z and rh axes. The third axis is the pressure
at each point in the pump in pascal.
In position (1) in Fig. 15, the meshing line is on the verge of
opening the first chamber to the discharge. In position (2), the
chamber is partially open and the pressure is building up. Because
the screw is double threaded, two chambers open to discharge in
every revolution as shown in positions (3) and (4).
Figure 16 presents a comparison between dynamic pressure
measurements from Ref. [5] and predictions from the model. The
dynamic pressure sensor is mounted in the housing at the position
of the last chamber (92 mm). During a complete rotation, the sensor
sees the discharge pressure, chamber 3 pressure, and values in
between on the screw land. Since the sensor is fixed in the housing
opposite to the meshing line for a spinning observer, the model
predictions is a record of the pressure values at a point following
the meshing line but lagging 180 deg.
The spectrum of the dynamic pressure is shown in Fig. 17. The
prediction matches the reported test data closely. Because the
pump is double threaded, the first harmonic appears at twice the
running speed, the other harmonics are at four, six, eight, and ten
times the running speed. The pressure at the zero frequency is the
static unbalanced hydraulic pressure.
Dynamic and Static Hydraulic Forces Acting on Screw
Rotor
Having established and validated a model that predicts the
dynamic pressure field as a function of the rotor angle, the next
step is to evaluate the forces acting on the rotor.
At any given axial location (z) and a screw rotational angle (/),
the forces can be calculated from the circumferential pressure
Fig. 14 Unwrapped screw geometry
062502-8 / Vol. 135, JUNE 2013 Transactions of the ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
Fig. 15 Two dimensional pressure field
Fig. 16 Dynamic pressure (GVF575%, z592mm) test data from Ref. [5]
Fig. 17 Dynamic pressure (GVF575%, z592mm) test data from Ref. [5]
Journal of Engineering for Gas Turbines and Power JUNE 2013, Vol. 135 / 062502-9
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
distribution and resolved in two perpendicular axes by the summation
in Eq. (13). In Fig. 14 the rotational angle (/) is the angle

between the meshing line, and the screw thread terminal edge,
while (h) is the angle around the rotor circumference measured
from an arbitrary plain that cuts the screw surface axially to
unwrap the geometry.
Fxz; /
Xh2p
i0
Pz; hdAhcosh
Fyz; /
Xh2p
h0
Pz; hdAhsinh
(13)
dA is equal to (dzRdh) for an angular point on a screw land and
(dzrdh) for an angular point on a chamber. The directions of the X
and y axes are shown in Fig. 14 to the right.
A schematic diagram of the axisymmetric structure model of a
rotor from Vetter s pump [5] is shown in Fig. 18. The bearing s
stiffness and damping are 6.11107 N/m and 525 N-s/m, respectively.
The mass and stiffness diameters of the screw section are
86.6mm and 68 mm, respectively. A dynamic finite element
model based on Timoshenko beam theory as in Ref. [23] gives the
rotor-bearing system s mass (M), damping (C), gyroscopic (G),
and stiffness (K) matrices governed by the matrix ODE in Eq.
(14). In this model, the damping in the system is only from the
bearings.
Mq C XGq_ Kq fqt (14)
where q is the vector of the system s translational and angular
degrees of freedom as defined in Ref. [23]. For external forces
with discrete frequencies components x1, x2, xn, the response
to each frequency component is given by
Q xj
 
x2
jM jxjXG C K
h i1
Fqxj (15)
From linearity, the total response is the summation of the
responses at each frequency. The maximum and minimum amplitudes
are calculated from a time series summation response of the
asynchronous frequencies.
To apply the hydraulic nonsynchronous forces on the rotor, the
120mm long screw is sliced into four sections as shown in Fig.
18. Each section is 30mm long. The forces acting on each section
are summed and taken to be applied at the center of the section.
The variation of forces at four axial locations in two complete
revolutions and the forces spectrum are shown in Fig. 19 for the x
direction and in Fig. 20 for the y direction.
The force in the direction of the common axis of the two rotors
(y axis) at the axial location closest to discharge (z105 mm) is
in opposite direction of the forces at the other three axial locations.
Therefore, the net effect of the y forces is almost cancelled
out. Whereas, the forces in the direction normal to the common
axis of the rotor (x axis) are all negative; therefore, there is a net
unbalanced hydraulic (both static and dynamic) forces in this
direction. These results are in line with the published static deflection
test results in Ref. [3].
The forces at zero frequency are applied as static forces, and
the deflection is calculated by
Qstatic K1Fx 0 (16)

The nonzero frequency forces are applied as nonsynchronous


stationary forces, and the response is calculated from Eq. [15].
The response spectrum in the x direction is shown in Fig. 21. The
response in the y direction was negligible compared to the
response in the x direction.
The static deflection at the discharge side of the screw is
1.6910 4 m (90% of the circumferential clearance), the maximum
amplitude of the time domain of the nonsynchronous
Fig. 18 Axisymmetric structure model for double thread twin screw pump from Ref.
[5]
Fig. 19 Hydraulic forces and hydraulic forces spectrum in x
direction at four axial locations ((z) in mm)
062502-10 / Vol. 135, JUNE 2013 Transactions of the ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
response is 910 6 m (5% of the circumferential clearance).
Unlike the rotor in Fig. 1, the rotors in the pump in Ref. [5] are
double thread. Therefore, the cross section is dynamically balanced,
and no synchronous response at the running speed is predicted.
In general, response to external unbalance can be
estimated similar to conventional rotating machinery.
The magnitude of the first harmonic of the hydraulic forces at
twice the running speed is around three times as high as the third
harmonic at six times the running speed. However, the response
due to the third harmonic is twice as much as the response due to
the first harmonic, since the higher harmonic is closer to the first
critical speed. This result shows that although the machine is operating
subcritical, the harmonics of the hydraulic forces can cause
excitation near the first natural frequency.
Conclusion
The model presented is a reasonable starting point for predicting
the steady-state vibration behavior of twin screw pumps. The
FE results showing that the screw geometry can be represented by
axisymmetric elements for both the lateral and torsional directions
are convenient. The hydraulic models available in the literature do
a decent job to predict chamber pressure in steady state operation.
The pressure drop between the chambers circumferential clearances
is taken to be linear similar to annular seals. The bridge
between the structural and hydraulic models is through expressing
the pressure field around the screws as a function of the rotational
angle. The results show oscillating hydraulic forces in multiples
of running speed that have the possibility of exciting the rotor s
natural frequency.
While the model includes the effect of the static rotor deflection
on the steady state flow between chambers (the circumferential
clearances are dependent on the static deflected rotor profile), it
does not capture the change in the dynamic pressure field due to
the rotor s lateral displacement. Instead it only predicts the unbalanced
dynamic pressure field due to the rotation of the helical
thread. The logical next step to improve the model would be to
relate the pressure field variation around screw lands to rotor perturbation
through stiffness, damping, and possibly added mass
terms, in addition to the unbalanced dynamic hydraulic forces presented
here. There is wide room for other improvement; however,
the model would be of little value if not validated by test results.
A rotordynamic testing program for twin screw pump should
include:
(1) dynamic pressure measurements preferably at multiple circumferential
points for each axial location along the screw
length

(2) dynamic displacement measurements at the center of the


rotor
(3) external housing shakers to excite the rotor natural frequency
in wet operation to estimate the damping provided
by the working fluid similar to the work by Kanki et al.
[24]
These measurements will help to validate the model and guide
the direction of its improvement. For example, estimated damping
and added mass terms in wet operation will give an insight of
whether a stiffness, damping, and added mass short seal terms can
be used to model the circumferential clearance.
Our hope is that this work opens a discussion involving users
and designers to elevate screw pump rotordynamic analysis capability
to par with the conventional centrifugal turbo machines.
Besides twin screw pumps, the rotordynamics of screw compressors
have not been adequately studied either. Screw compressors
have more complicated geometry than screw pumps as well as
pronounced acoustic effects. However, a formal, validated rotordynamic
model of screw pumps will inevitably draw attention to
the need for a rotordynamic model of screw compressors.
Nomenclature
Agap the gap area between the screw and liner opening to discharge
(m2)
At effective leakage area in the tip clearance space (m2)
dA incremental rectangular area on screw surface (m2)
C damping matrix of the FE rotor/bearing system
dh the hydraulic diameter (twice the gap height) (m)
d/D ratio of the orifice to pipe diameter
F(xi) Fourier transform of the hydraulic forces
f Blasius friction coefficient
fq external forces acting on the rotor/bearing system (N)
ft empirical constant to account for leakage paths other
than the circumferential clearance
G gyroscopic matrix of the FE rotor/bearing system
h screw thread pitch length (m)
K stiffness matrix of the FE rotor/bearing system
ke the entry loss coefficient for the circumferential
clearance
l the length of the circumferential clearance leakage path
across the screw land (m)
M mass matrix of the FE rotor/bearing system
Pd discharge pressure (Pa)
Pn pressure in the last chamber (Pa)
pi pressure in chamber number i (Pa)
Pi,prv pressure in chamber number i from previous iteration
(Pa)
Qi1!i circumferential slip flow between chambers i1 and i
(m3/s)
Qgap the backflow from of the gap area between the screw
and the liner (m3/s)
Fig. 20 Hydraulic forces and hydraulic forces spectrum in y
direction at four axial locations ((z) in mm)
Fig. 21 Nonsynchronous dynamic response in the x direction
Journal of Engineering for Gas Turbines and Power JUNE 2013, Vol. 135 / 062502-1
1
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms
Q(xi) Fourier Transform of the response of the rotor-bearing
system FE degrees of freedom
R, r screw outer, and inner radii (m)

VK the chamber volume (m3)


Vg,i the gas volume in chamber number i (m3)
VL,i the liquid volume in chamber number i (m3)
Vlkg leakage velocity across the circumferential screw land
(m/s)
Vin volume of liquid flow leaking in the chamber from the
chamber downstream (m3)
Vout volume of liquid flow leaking out of the chamber to the
chamber upstream (m3)
z displacement coordinate of the axis across the axial
direction of the screw
b bulk modulus of elasticity of the fluid (N/m2)
b 0
effective bulk modulus of elasticity of the fluid and gas
mixture (N/m2)
c ratio of specific heats for the gas
h angular coordinate of the polar axis of the rotor
ql fluid density (kg/m2)
/ the angle of the meshing line defined in the polar coordinate
of the rotor (rad)
x discreet frequency (asynchronous) (rad/s)
X running speed (rad/s)
Subscripts
d discharge
i chamber number i
n last chamber opening to discharge
s suction
Abbreviation
GVF gas volume fraction Vg=Vg Vfl
 
References
[1] Neumann, W., 1991, Efficient Multiphase Pump Station for Onshore Application
and Prospects for Offshore Application, Proceedings of the 8th International
Pump User Symposium, Turbomachinery Laboratory, Texas A&M
University, College Station, TX, March 5 7, pp. 43 48.
[2] Shippen, M. E., and Scott, S. L., 2002, Multiphase Pumping as an Alternative
to Conventional Separation, Pumping and Compression, Proceedings of the
34th PSIG Conference, SPE Reprint, Portland, OR, October 23 25, Vol. 58.
[3] Vetter, G., and Wincek, M., 1993, Performance Prediction of Twin-Screw Pumps
for Two-Phase Gas/Liquid Flow, Proceedings of the Fluid Engineering Conference
(FED), Pumping Machinery, ASME, Washington, D.C., Vol. 154, pp. 331 340.
[4] Prang, A. J., and Cooper, P., 2004, Enhanced Multiphase FLow Predictions in
Twin-Screw Pumps, Proceedings of the 21st International Pump User Symposium,
Baltimore, MD, March 8 11, Turbomachinery Laboratory, Texas A&M
University, College Station, TX, pp. 69 76.
[5] Vetter, G., Wirth, W., Korner, H., and Pregler, S., 2000, Multiphase Pumping
With Twin-Screw Pumps Understand and Model Hydrodynamics and Hydroabrasive
Wear, Proceedings of the 17th International Pump User Symposium,
Houston, TX, March 6 9, Turbomachinery Laboratory, Texas A&M University,
College Station, TX, pp. 153 169.
[6] Rausch, T., Vauth, Th., Brandt, J.-U., and Mewes, D., 2004, A Model for the
Delivering Characteristic of Multiphase Pumps, Proceedings of the 4th Annual
North American Conference on Multiphase Technology, Banff, Canada, June
3 4, pp. 313 325.
[7] Martin, A. M., 2003, Multiphase Twin-Screw Pump Modeling for the Oil and
Gas Industry, Ph.D. thesis, Petroleum Engineering Department, Texas A&M
University, College Station, TX.
[8] XU, J., 2008, Modeling of Wet Gas Compression in Twin-Screw Multiphase
Pump, Ph.D. thesis, Petroleum Engineering Department, Texas A&M University,
College Station, TX.

[9] Chan, E., 2006, Wet-Gas Compression in Twin-Screw Multiphase Pumps


M.S. thesis, Petroleum Engineering Department, Texas A&M University, College
Station, TX.
[10] Rabiger, K., Maksoud, T. M. A., Ward, J., and Hausmann, G., 2008,
Theoretical and Experimental Analysis of a Multiphase Screw Pump, Handling
Gas Liquid Mixtures With Very High Gas Volume Fractions, Exp.
Therm. Fluid Sci., 32, pp. 1694 1701.
[11] Rabiger, K., 2009, Fluid Dynamic and Thermodynamic Behaviour of Multiphase
Screw Pumps Handling Gas-Liquid Mixtures With Very High Gas Volume
Fractions, Ph.D. thesis, University of Glamorgan, Treforest, UK.
[12] Hamelberg, F. W., 1966, Lauferkrafte bei Schraubenpumpen, Ph.D. thesis,
Technical University of Hanover, Hanover, Germany.
[13] Stosic, N., Smith, I., and Kovacevic, A., 2005, Screw Compressors Mathemati
cal
Modeling and Performance Calculation, Springer, New York.
[14] Stosic, N., Smith, I. K., Kovacevic, A., and Mujic, E., 2006, Vacuum and Mul
tiphase
Screw Pump Rotor Profiles and Their Calculation Models, VDI-Ber.,
1932, pp. 53 68.
[15] Rao, J. S., Shiau, T., and Chang, J., 1998, Theoretical Analysis of Lateral
Response Due to Torsional Excitation of Geared Rotors, Mech. Mach. Theory,
33(6), pp. 761 783.
[16] Lee, A. S., Ha, J. W., and Choi, D.-H., 2003, Coupled Lateral and Torsional
Vibration Characteristics of a Speed Increasing Geared Rotor-Bearing System,
J. Sound Vib., 263, pp. 725 742.
[17] Childs, D., 1993, Turbomachinery Rotordynamics: Phenomena, Modeling &
Analysis, John Wiley & Sons, New York.
[18] Singh, A., 2003, Modeling Twin-Screw Multiphase Pump Performance During
Periods of High Gas Volume Fraction, M.S. thesis, Petroleum Engineering
Department, Texas A&M University, College Station, TX.
[19] Manring, N. D., 2000, The Discharge Flow Ripple of an Axial-Piston SwashPlate Type Hydrostatic Pump, ASME J Dyn. Syst., Meas., Control, 122, pp.
263 268.
[20] Cho, B. H., Lee, H. W., and Oh J. S., 2000, Estimation Technique of Air Cont
ent
in Automatic Transmission Fluid by Measuring Effective Bulk Modulus,
Proceedings of the FISITA World Automotive Congress, Seoul, Korea, June
12 15.
[21] Tullis, J., 1989, Hydraulics of Pipeline: Pumps, Valves, Cavitation, Transi
ents,
John Wiley & Sons, New York.
[22] Martino, G., Fontana, N., and Giugni, M., 2008, Transient Flow Caused
by Air Expulsion Through an Orifice, J. Hydraul. Eng., 134(9), pp.
1395 1399.
[23] Nelson, H. D., 1980, A Finite Rotating Shaft Element Using Timoshenko
Beam Theory, ASME J. Mech. Des., 102(4), pp. 793 803.
[24] Kanki, H., Fujii, H., Hizume, A., Ichimura, T., and Yamamoto, T., 1986,
Solving Nonsynchronous Vibration Problems of Large Rotating Machineries
by Exciting Test in Actual Operating Condition, Proceedings of the IFToMM
International Conference on Rotordynamics, Tokyo, Japan, September 14 17,
pp. 221 226.
062502-12 / Vol. 135, JUNE 2013 Transactions of the ASME
Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 03/0
3/2015 Terms of Use: http://asme.org/terms

You might also like