Sie sind auf Seite 1von 29

Internal Ballistics of PCP Airguns

Domingo Tavella, Ph.D.

Abstract
This work presents a quasi-stationary, inviscid gas dynamic framework for the analy-
sis of pre-compressed (PCP) airguns. The framework methodology, despite its simplicity,
clarifies and explains some of the more intriguing aspects of PCP airgun internal ballis-
tics, such as the presence and shot-count location of optimal reservoir pressure, and fac-
tors that determine the gun performance. A case study demonstrates the potential of the
methodology.

1.0 Background
Air guns, or more generally, gas guns, extract energy from a compressed gas to propel a
projectile. There are two basic architectures to build an airgun. A spring piston airgun
relies on a compressed spring (metallic or pneumatic), which, when the gun is fired, com-
presses the air that drives the projectile down the barrel. I discussed the internal ballistics
of spring piston airguns and associated issues in previous work [1, 2, 3].

The other fundamental architecture, which concerns us here, uses a high-pressure reser-
voir as a source of high-pressure air. In small, portable airguns, the reservoir is part of the
gun. In large airguns, such as artillery pieces, the reservoir can be physically separate from
the gun itself.

Small PCP airguns have become extremely popular in the last few years, with their inter-
nals having evolved to a high degree of sophistication, and their accuracy rivaling that of
high quality conventional firearms. Historically, PCP airguns have been around for a long
time, and while their most common application is in sport shooting (all modern olympic
air rifles are of the PCP type,) their concept also found application in warfare and hyper-
velocity research. In the late 1800’s, for example, the city of San Francisco (US) was
defended by a battery of PCP air cannon, capable of launching 500 lb projectiles over a
mile out to sea [4], and PCP mortars were used during WWI [5]. While modern artillery
displaced PCP artillery long ago, sophisticated variations on gas guns are used today to
simulate the effects of micro-meteorite impacts, where extremely high velocities must be
achieved [6].

The firing cycle of a standard PCP airgun begins when the trigger sear releases a hammer
(or striker), which is held against a compressed spring1. The hammer accelerates under the
force of the spring, and after travelling a short distance it strikes a valve that communi-
cates with a high-pressure source (a reservoir or a regulator). Upon being struck by the

1. Some airguns use solenoids to open the reservoir valve.

Internal Ballistics of PCP Airguns August 16, 2018 1


hammer, the valve briefly opens, allowing a blast of high-pressure air to be routed to the
breech, where pressure builds up and propels the pellet down the barrel. During the few
milliseconds the firing cycle lasts, a number of enthalpy and mass transfer processes take
place, whose precise characterization is essential to capture the physics involved.

Figure 1 shows the simplest model configuration with the components you need to make
sense of the internal ballistics of a conventional, non-regulated1 PCP airgun.

FIGURE 1. PCP airgun layout for analysis.

An interesting sequence of events follow the trigger release, whose timing depends on the
particulars of the gun. In the following description, the timing of these events is approxi-
mate and corresponds to the case study discussed in detail below. The striker takes about
five milliseconds to reach the valve stem. When the hammer strikes the valve, there is an
exchange of momentum between the hammer and the valve, which, if of adequate inten-
sity, causes the valve to open. The valve reaches maximum lift in less than half a millisec-
ond after being struck by the hammer, and closes about half of a millisecond later. During
the time the valve remains open, a blast of high-pressure air discharges into a space that
connects to the transfer port - I will refer to this space as “transfer channel”, “interstitial
space”, or “transfer plenum” (notice that “transfer port” refers to the opening connecting
this space to the barrel.) The transfer channel plays an important role in the gun perfor-
mance. The transfer channel accumulates air at the same time as it releases it through the
transfer port into the barrel. The pellet barely moves during the time the valve reaches
maximum lift, and by the time the valve closes, the pellet has moved less than 10 cm down
the barrel. After the valve closes, air accumulated into the transfer channel continues to
flow into the barrel, and the hammer bounces backwards. After about eight milliseconds
after hammer release, the pellet leaves the barrel, with the hammer about halfway in its
backward bounce. About 7 milliseconds later, the hammer bounces forward once again,
striking the valve stem for the second time, triggering further release of air. By now, the
pellet is on its way to the target. Keep in mind that both the timing of these events, and the
amount of air released at each hammer strike are dependent on the problem parameters.

1. A regulated PCP airgun incorporates a regulator that supplies high-pressure air at a constant pressure.
The analysis in this study applies without modification to a regulated gun.

Internal Ballistics of PCP Airguns August 16, 2018 2


What happens within a PCP airgun involves a mix of elasticity, thermodynamic, and gas
dynamics elements. The components of the gun interact in such a way that an optimal con-
figuration exist for which the gun performance is optimal. Optimal performance is not
uniquely defined. A very common optimization task in a PCP airgun is to determine the
reservoir pressure for which, all else remaining constant, there is a maximum transfer of
energy to the pellet.

The objective of this study is to develop a simple model to explore some of the most
important issues of a simple PCP airgun, and to formulate a framework that can accommo-
date improvements of the model component parts at a later time.

A small number of publications in the open literature germane to PCP airguns deal with
simple case studies aimed at undergraduate college students [7, 8, 9]. Horak et al [10] dis-
cuss the modeling of airsoft guns, which have some issues in common with the PCP air-
guns of interest to us here. A number of publications deal with hyper-velocity gas guns,
which share some of the modeling issues of PCP airguns [11, 12].

2.0 Analysis
The analysis will work on the stylized configuration sketched in Figure 1. Configurations
can be significantly more complex, but this simple layout is sufficient to capture the fun-
damental physics involved. This simple layout has at least ten basic elements which define
the gun performance: 1) the hammer (or striker), 2) the hammer spring, 3) the valve and its
stem, 4) the valve spring, 5) the high-pressure reservoir, 6) the transfer plenum or transfer
channel, 7) the transfer port, 8) barrel length, 9) projectile mass, and 10) projectile caliber.
The fact that there are at least ten elements, and given that there are only three fundamen-
tal dimensions - length, time, and force (or mass) - immediately tells you that the behavior
of a PCP airgun is a problem with at least seven independent dimensions. This also means
that, if you do things right, you should end up with a system of seven independent equa-
tions describing the problem.

2.1 Hammer and valve impact


There are three types of collisions you must consider in a PCP mechanism. The most
important one is the impact of the hammer against the valve stem when the valve is ini-
tially closed. Another is the collision between hammer and valve that might happen while
the valve is open. Yet another is the impact the valve makes against its seat when it closes.
These are all two-body type collisions. Theoretically, there is the possibility of a three-
body collision if the valve were to be hit by the hammer as the valve makes contact with
its seat during closure. I will assume this last type of collision is extremely unlikely and
will ignore it. Let’s consider the hammer-valve impact when the valve is first hit by the
hammer.

To precisely determine whether the hammer will open the valve when it hits the valve
stem you have to solve a complex open two-body impact problem. A collision problem is
“open” when external forces interfere with the momentum exchange between the colliding

Internal Ballistics of PCP Airguns August 16, 2018 3


bodies. You can solve the impact problem in a PCP airgun at several levels of complexity,
depending on the time scales involved. If the time scale of the collision is comparable to
the valve dwell time scale, you must solve the collision problem in full, accounting for
what happens during the collision. Hertz [13] laid out the elasticity framework for dealing
with colliding bodies when you need full resolution of the collision process. Analytical
solutions exist for relatively simple bodies, such as spheres, bars, and planes [14]. For
more complex colliding bodies, such as PCP hammers and valves, detailed resolution of
the collision process requires a numerical technique such as finite elements. Since the
focus here is on simplicity and computational economy, it is best to find a simpler alterna-
tive.

If the collision time scale is much shorter than the time the valve remains open, you can
simplify the problem by assuming the collision occurs instantaneously. In the case of a
PCP hammer-valve impact, the valve must overcome some deflection of its seating area,
and this means that the impact of the hammer with the valve stem will happen as an open
system, subject to external momentum transfer. Elementary collision theory, as is usually
presented in elementary Physics texts, assumes the colliding bodies form a closed system,
subject to no external forces.

I will assume the collision time is short enough that it can be neglected compared with the
dwell time of the valve. Without entering into the mathematical details, this assumption is
justified by comparing computed valve dwell times with characteristic collision times of
simple bodies (such as spheres of appropriate size and mass) using Hertz’ theory. If you
replace the hammer and valve with spheres of diameter and mass comparable to the diam-
eter and mass of the hammer and valve stem, Hertz theory gives you collision times of the
order of 50 microseconds, or 0.05 millisecond. Typical dwell times for the configurations
studied here are about one millisecond, or twenty times longer. To determine how good
this assumption really is would require empirical verification with sophisticated measure-
ment equipment. One of the consequences of the instantaneous collision assumption is the
possibility of multiple collisions between the hammer and the valve stem while the valve
is open.

The instantaneous collision assumption when there is an external source of impulse


amounts to a separation of the momentum exchange between hammer and valve during
collision, and the momentum exchange between the valve and the closing forces acting on
the valve (the valve spring and the reservoir pressure). In our case, the external impulse
comes from the resistance of the valve to be lifted against the reservoir pressure as the
valve separates from its deflected seat. In this approximation, the bouncing velocities of
the hammer and valve result from elementary collision theory, and whether the valve can
overcome the forces that keep is the outcome of a-posteriority verification.

If v h is the hammer velocity when it first makes contact with the valve stem, m h is the
hammer mass, and m v is the mass of the valve, the hammer and valve velocities after
impact are given by,

Internal Ballistics of PCP Airguns August 16, 2018 4


+ mh – mv e
v h = v h ---------------------- (EQ 1)
mh + mv

+ mh  1 + e 
v v = v h ------------------------ (EQ 2)
mh + mv

where e is the restitution coefficient of the hammer and valve metal pair (typical values of
e for steel are between 0.7 and 0.85.) The restitution coefficient represents the fraction of
momentum that is restored to the colliding bodies as they bounce away from each other.

In this simplified approach, in order for the valve to open, the energy of the valve after
impact with the hammer must be greater than the energy used in overcoming the forces
that keep the valve closed. Mathematically, this means the valve velocity calculated with
EQ 2 must be corrected by the energy consumed in restoring the valve seat to its unloaded
shape,

2 +
+ +2  Av p0 
vv  vv –  ------------------- (EQ 3)
m v ES

where S is the valve seating area,  is a characteristic thickness of the valve seating area,
p 0 is the initial reservoir pressure, and E is Young’s elasticity modulus of the valve seat-
+
ing material. The operator   returns the positive value of its argument if the argument
is greater than zero, or zero otherwise. The thickness  and surface S are functions of the
valve and seat geometry, and it is best to treat these quantities as calibration parameters.
EQ 3 ignores the effect of the initial transfer plenum pressure on the valve poppet, which
is insignificant when compared with the initial reservoir pressure.

+
Once you know the minimum value of v v required to open the valve, you can use EQ 1 to
calculate the minimum hammer velocity upon valve contact, and with this it is trivial to
compute the minimum hammer spring strength required to open the valve.

A different situation occurs if the hammer hits the valve stem while the valve is open. In
this case, both the hammer and the valve are moving before collision. Momentum preser-
vation yields,

+ mh – mv e mv  1 + e 
v h = v h ---------------------- + v v ----------------------- (EQ 4)
mh + mv mh + mv

+ mh  1 + e  mh – mv e
v v = v h ------------------------ + v v ---------------------- (EQ 5)
mh + mv mh + mv

As mentioned earlier, these equations ignore momentum interference with external forces
- this is only a working assumption, subject to experimental verification. The intuition

Internal Ballistics of PCP Airguns August 16, 2018 5


behind this assumption is that if the valve is hit while open, the valve “break-open” time
scale is absent, and simple collision theory will suffice.

The third type of collision is when the valve returns to its seat and closes. It is theoretically
possible that the valve may bounce when it hits its seating area, but I will assume this
doesn’t happen. In other words, I will assume that if the valve re-opens after closing it is
because it was struck again by the hammer.

Figure 2 shows the hammer and valve coordinates and spring pre-compressions with the
hammer in the cocked position. Notice that the physical dimensions of the hammer are
irrelevant to the analysis. To simplify the analysis I will idealize the hammer as a thin
sliver of mass m h .

FIGURE 2. Hammer and valve coordinates and spring pre-compressions - L h is the spring
free length, L v is the initial distance, x h is the hammer coordinate, and x v is the valve
coordinate, all measured from the datum line, defined as the position of the spring base for
zero pre-compression adjustment.

As the figure shows, the initial hammer spring compression is the sum of three compo-
nents, a pre-compression adjustable by the user,  apc , a set (or default) pre-compression
determined by the free length of the spring,  dpc , and the cocking stroke,  cs . Ignoring
friction forces and assuming a perfect mass-less spring with constant k h , the hammer
velocity at the time of first contact with the valve stem is,

kh 2 2
vh  tv  = ------    apc +  dpc +  cs  –   apc +  dpc   (EQ 6)
mh

If the hammer is detached from the spring, the hammer equations of motion are,

Internal Ballistics of PCP Airguns August 16, 2018 6


2
d xh kh
= ------   apc +  dpc + L v – x h  x h  x v +  apc
dt
2 mh
(EQ 7)
2
d xh
= 0 x h  x v +  apc
2
dt

where k h is the hammer spring constant. These equations apply in-between strikes, with
initial values consistent with EQ 1 to 4. In practice, since the valve lift is very much
smaller than either the hammer travel or the adjustable pre-compression, it won’t make
much difference if you treat the hammer as being attached to the spring.

The motion of the valve follows a similar equation of motion between strikes,
2
d xv 1
= – ------  F + k v   vpc + x v – L v   , (EQ 8)
dt
2 mv

where m v is the mass of the valve and its stem,  vpc is the valve pre-compression, and F
is the net force acting on the valve poppet. The definition of F plays a critical role in the
mass and energy transfer from the reservoir into the transfer channel. The simplest defini-
tion F is as follows,

F = Av  pp – p0  – As pp (EQ 9)

where A v is the area of the valve poppet exposed to the reservoir pressure, A s is the cross-
sectional area of the valve stem, p p is the pressure in the transfer plenum, and p 0 is the
pressure in the reservoir. EQ 9 is an approximation; a precise definition of F requires
detailed resolution of the flow field in the neighborhood of the poppet. This, and the
dynamics within the transfer channel, are the two most important issues to be addressed
for more accurate computations.

Notice that while it is very simple to solve the hammer equation of motion analytically
between strikes, you have to solve the equation of the valve numerically, since the forcing
term F isn’t known analytically.

Internal Ballistics of PCP Airguns August 16, 2018 7


2.2 Mass flow transfer
Ejection from the reservoir into the transfer channel. With the critical pressure ratio,
 -
----------
–1
p =  -----------
* 2
, where  is the ratio of specific heats, the discharge into the plenum is
  + 1
pp *
supercritical if -----  p . The supercritical mass flow coefficient into the plenum is,
p0

-----------
+1
–1
  -----------
2
K sup in = (EQ 10)
+1

and the associated mass flow rate,

K sup in
q· sup in = ----------------- p 0 (EQ 11)
RT 0

where T 0 is the temperature in the reservoir. Notice the linear dependence on the reservoir
total pressure.

pp *
Subcritical injection into the transfer plenum happens when -----  p , and the associated
p0
subcritical mass flow coefficient and mass flow rates are,

2 –1
p p   p p  
--- -----------
K sub in =
2  
----------- ----- 1 – ----- (EQ 12)
 – 1  p 0   p 0 
 

K sub in
q· sub in = ----------------- p 0 (EQ 13)
RT 0

With q· in q· out denoting either the supercritical or the subcritical mass flow rate, the gas
mass per unit time injected into the plenum is,

dm p in
= C in A in q· in (EQ 14)
dt

where C in is a discharge correction coefficient, and A in is the area the valve clears to
allow the high pressure gas into the plenum. In the configuration shown in Figure 1, this
area is a linear function of the valve displacement, but in general it is possible to contour

Internal Ballistics of PCP Airguns August 16, 2018 8


the valve and its seat in such a manner that a specific dependence on valve displacement
obtains. For the case in Figure 1, therefore,

A in = 2  x v – L v  A v x v  L v (EQ 15)

Ejection from the transfer plenum into the barrel. Unlike the mass transfer into the
transfer plenum, where the reservoir stagnation values are known unequivocally, the mass
outflow is contingent on calculation assumptions. The assumption that you can neglect the
time change of volumetric kinetic energy is more of an issue in the transfer plenum, with
its small volume and two openings, than in the reservoir. Despite this potential source of
controversy, I will assume you can neglect the volumetric kinetic energy time changes in
the plenum in the energy balance. This amounts to saying that the static pressure in the
plenum, p p , is numerically equivalent to the stagnation pressure in the plenum. With this
assumption, and denoting the pressure at the breech by p breech , you can adapt the previ-
ous equations to the transfer port mass outflow.

p breech *
The supercritical mass outflow through the transfer port occurs when ----------------  p , and the
pp
associated transfer coefficient and mass flow rate are

-----------
+1
–1
  -----------
2
K sup out = (EQ 16)
+1

K sup out
q· sup out = ------------------- p p (EQ 17)
RT p

p breech *
and for the subcritical case, ----------------  p ,
pp

2 –1
p b   p breech  
--- -----------
K sub out =
2  
----------- ----- 1 – ---------------- (EQ 18)
 – 1  p p   pp  
 

K sub out
q· sub out = ------------------- p p (EQ 19)
RT p

The gas mass leaving the transfer plenum per unit time is,

dm p out
= C tp A tp q· out (EQ 20)
dt

Internal Ballistics of PCP Airguns August 16, 2018 9


were C tp is a discharge correction coefficient for the transfer port, and A tp is the area of
the transfer port.

Discharge correction coefficients. The quantities C in and C tp reflect two aspects of the
flow that interfere with discharge - viscous effects and vena contracta effects. In the case
study I discuss below I ignored viscous effects, not because they are negligible, but
because they can be viewed as a calibration issue. I took into account the impact of com-
pressibility on the vena contracta effect, however, since this is consistent with an inviscid
framework [15].

2.3 Transfer plenum conservation equations

With  p the gas density in the plenum and V p the plenum volume, mass conservation is,

d p 1 dm dm p out
= ------  p in – (EQ 21)
dt Vp d t dt 

This equation assumes the plenum volume doesn’t change during firing. This isn’t strictly
true, since the valve displacement enlarges the intersticial volume. However, this assump-
tion is justified when you consider that the volume subtended by the valve lift is much
smaller than the volume of the transfer plenum (see [1] for detailed derivation of the con-
servation equations.)

Under adiabatic conditions, the energy equation in terms of the transfer plenum tempera-
ture is,

dT p 1 dm p in  dm p out 
T in – T p + -------- v in – T out – T p + -------- v out
1 2 1 2
= ------ (EQ 22)
dt mp d t  2c v  dt  2c v 

where T in T out are the discharge temperatures of the gas entering and leaving the plenum,
respectively, and v in v out are the corresponding discharge velocities. As was the case with
the mass conservation equation, this equation also assumes the volume of the transfer ple-
num is constant. This assumption has an impact on the energy content in the plenum in
that it ignores the force times displacement component of energy. The basis for ignoring
this component of external work is the same as for the mass conservation equation - the
valve displacement is very small compared with other dimensions in the plenum. A more
important assumption in deriving the energy equation, however, is that I ignore the kinetic
energy time in the transfer plenum.

The discharge quantities in the energy equation result from assuming locally isentropic
expansions at the valve opening and transfer port planes. The discharge pressure from the
valve into the transfer plenum, p d in , is, in the supercritical case,

Internal Ballistics of PCP Airguns August 16, 2018 10


*
p d in = p 0 p (EQ 23)

and in the subcritical case,

p d in = p p (EQ 24)

With  0 the gas density in the reservoir, the discharge velocity and temperature from the
reservoir into the transfer plenum are,
–-----1-
p d in  p
v in = K in  ----------- ----0- (EQ 25)
 p0  0

----------
– 1-
p d in 
T in = T 0  ----------- (EQ 26)
 p0 

where K in stands for either K sub in or K sup in , depending on whether the discharge
regime is subcritical or supercritical.

The discharge quantities from the plenum into the barrel through the transfer port are sim-
ilarly derived, but using the gas in the transfer plenum as the reference stagnation state.
The discharge pressure through the transfer port is, therefore,
*
p d out = p p p (EQ 27)

in the critical case, and

p d out = p breech , (EQ 28)

in the subcritical case. These last two equations provide the linkage between the thermo-
dynamics of the reservoir and the intersticial space, and the ballistics in the barrel.

The discharge velocity and temperature, for the subcritical case, result from re-interpreting
the stagnation conditions in EQ 25 and 26,
–-----1-
p d out  p
v out = K out  -------------- ----p- (EQ 29)
 pp  p

----------
– 1-
p d out 
T out = T p  -------------- (EQ 30)
pp

where K out represents either K sub out or K sub out , depending on the discharge regime.
Working with this framework, where you switch the stagnation conditions from the reser-

Internal Ballistics of PCP Airguns August 16, 2018 11


voir to the transfer plenum, requires a clear understanding that invariance of stagnation
temperature is only valid in steady state. The energy transfers from reservoir to plenum to
barrel is not a steady state problem - keeping this in mind is important for proper interpre-
tation of the results.

2.3.1 Barrel ballistics

Conservation equations. The conservation equations in the transfer plenum must connect
with the conservation equations in the barrel behind the projectile. This connection hap-
pens through the discharge pressure from the transfer plenum in the barrel, and from the
corresponding mass flow rate through the transfer port. The next figure illustrates the
transfer port-barrel set up.

FIGURE 3. Communication of transfer plenum with barrel.

The mass flow into the barrel matches the mass outflow from the transfer plenum,

dm b dm p out
= (EQ 31)
dt dt

where m b is the gas mass behind the pellet. With A b the cross-sectional area of the bore,
 b the gas density in the bore behind the pellet, x b the rear-end position of the projectile,
measured from the back of the breech, and v b the projectile velocity, the mass conserva-
tion differential equation is,

d b 1 dm b
= ----------- – Ab vb b (EQ 32)
dt Ab xb d t

Here I introduce the important assumption that the gas density in the bore between the
breech and projectile is uniform. This assumption, whose adaptation to airguns I discuss in
the next section, is central to the internal ballistics of firearms. The initial value of x b is
the seating depth of the pellet in the barrel. This depth must be large enough to accommo-
date the transfer port.

The energy equation reflects the balance between enthalpy inflow through the transfer
port, energy expended pushing the projectile down the barrel, and the change of kinetic
energy in the bore behind the pellet (see [1] for detailed derivation),

Internal Ballistics of PCP Airguns August 16, 2018 12


dT b 1 dm b  Ab
T out – T breech + -------- v out – ------ p base v b – ---- d KE
1 2 1
= ------ (EQ 33)
dt mb d t  2c v  cv cv d t

where T breech is the temperature at the breech, p base is the pressure at the pellet base, and
KE is the kinetic energy of the gas behind the pellet.

The momentum equation closes the system and allows for the calculation of the pellet
position and velocity. With M the pellet mass, p nose the pressure on the nose of the pellet,
F break the break force required to get the pellet moving, and F the friction force, momen-
tum balance gives,

dv b F break
= 0 ; p base  p nose + --------------
-
dt Ab
(EQ 34)
dv b A b F break
= ------  p base – p nose  – F ; p base  p nose + --------------
-
dt M Ab

dx b
= vb (EQ 35)
dt

In general, the nose pressure is a function of the pellet acceleration as it expels the volume
of air in the bore between the pellet position and the muzzle. As I showed in my previous
work on spring-piston airguns [1], this component of inertia is very small and you can
safely neglect it. In addition to the friction force, there is a force that reflects the angular
acceleration of the projectile as it engages the rifling. In [1], I showed that this component
is also small and you can neglect it.

Lagrange gradient approximation (LGA). The LGA is a simple and effective way to
take into account the inertia of the gas behind the projectile in calculating both the acceler-
ation of the projectile in the bore and the breech pressure to which the transfer port is
exposed. The LGA is based on the assumption that you can posit the density behind the
projectile to be uniform along the bore. This assumption, when incorporated into the
momentum conservation equation, yields a closed-form expression for pressure variation
along the bore [16]. The LGA has been a standard assumption in internal ballistics of fire-
arms for a very long time, but until now it has never, to my knowledge, been extended to
airguns.

Except for very small calibers, the powder in firearm ammunition is comparable in weight
to the projectile itself. In fact, in large caliber artillery, the propellent weight can be signif-
icantly higher than the weight of the projectile. The reason why the LGA works extremely
well in firearms is that you can assume that the powder charge burns uniformly, and the
resultant gas can expand with a negligible density gradient as the bullet accelerates down
the bore. Airguns are different from firearms in that the high-pressure gas doesn’t emerge
all at once in a small volume, but rather it gets injected into the breech over a period of

Internal Ballistics of PCP Airguns August 16, 2018 13


time, and it is reasonable to assume that this injection over time creates a pressure gradient
down the bore. However, as will emerge from the computations in this work, in the case of
PCP airguns, much of the high-pressure blast of gas into the bore occurs early in the pel-
let’s travel down the bore. In this regard, PCP airguns are similar to firearms, and this
motivates the adaptation of the LGA to airguns.

Besides these plausibility arguments for adapting the LGA to airguns, there is computa-
tional evidence that when gases expand in tubes behind a sliding body, such as a projectile
of a piston, the density tends to remain relatively uniform [17].

In the case study I will discuss below, the mass of air typically injected behind the pellet is
of the order of 10% of the pellet mass, which suggests that including the gas mass behind
the pellet may make a difference worth capturing.

Notice that the gas inertia behind the pellet enters into the computations both via the
energy equation, through the term d KE , and in the momentum equation by inducing a
dt
longitudinal pressure gradient. This pressure gradient increases p breech and decreases
p base . The increment in breech pressure tends to lower the mass flow through the trans-
fer port, and the decrement in base pressure decreases the momentum transfer to the pro-
jectile. This suggests that the inertia of the gas behind the pellet influences the gun
performance in three connected and negative ways, which warrants investigating the
issue.

The standard LGA as applied to firearms assumes the combustion gas velocity at the
breech is zero. To adapt the LGA to airguns, I will modify this assumption by assigning a
value to the gas velocity at the breech in the amount,

1 dm b
v breech = ------------ (EQ 36)
Ab b d t

Assigning the velocity v breech to the gas at the breech is equivalent to assuming there is
instantaneous mixing of the gas emerging from the transfer port. The LGA also assumes
that the temperature (or internal energy) that results from solving the energy equation in
the barrel is an average temperature consistent with an average pressure and uniform den-
sity. This interpretation of the energy equation is central to the LGA applied to firearms,
where the reaction rate of the propellent is assumed to respond to the average temperature
and pressure.

Working with the inviscid compressible momentum equation, after some algebra you
obtain a linear distribution of the gas velocity between the breech and the projectile,

x
v  x  = v breech +  v b – v breech  ----- (EQ 37)
xb

Internal Ballistics of PCP Airguns August 16, 2018 14


With a linear distribution of velocity you get a linear pressure gradient, which you can
integrate between the breech and the base of the projectile to get a quadratic distribution of
pressure behind the pellet. With a bit more algebra, you can extract the pressure at the
breech and the pressure at the base of the projectile in terms of the average pressure and
the gas acceleration at the breech and at the pellet base (the gas velocity at the pellet base
equals the pellet velocity.) In the following two equations, p b is the average pressure
behind the pellet - this is the pressure that links, through the equation of state, the temper-
ature in EQ 33 with the uniform density behind the projectile.

m b 1 dv breech 1 dv b
p breech = p b + ------  --- + --- (EQ 38)
Ab  3 d t 6dt 

m b 1 dv breech 1 dv b
p base = p b + ------  --- + ---  (EQ 39)
Ab 6 d t 3 dt

dv b
Replacing from EQ 34, the projectile acceleration is,
dt

dv b 1 m dv breech
= ---------------------------- A b  p base – p nose  – ------g –F (EQ 40)
dt  mg  6 dt
M 1 + --------
 3M

In the standard form of the LGA as applied to firearms, the middle term in EQ 40 is miss-
ing and the propellent charge replaces the gas mass. Notice that the bullet accelerates as if
its mass had been raised by one third of the ratio of the gas mass behind the pellet to the
pellet mass, and as if the force applied to the pellet had decreased by one sixth the inertia
of the gas mass, computed as if the gas mass accelerated with the acceleration at the
dv breech
breech. The following expression for rounds up these formulas,
dt

2
dv breech 1  d m g ----
1 d b dm g
= ------------  – -  (EQ 41)
dt Ab b  d t2 b d t d t 

You can now evaluate p breech , and use this pressure to determine the transfer port flows.

Kinetic energy of the gas behind the pellet. The kinetic energy content of the gas
1 xb
behind the projectile is KE = --- A b  v dx ,
2
2 0

1 2
KE = --- A b  b  v breech  v breech + v b  + v b x b (EQ 42)
6

Internal Ballistics of PCP Airguns August 16, 2018 15


d
The most practical way to compute KE in the energy equation is by time-lagged finite
dt
differences.

3.0 Analysis summary


A set of seven coupled, mostly non-linear ordinary differential equations describe the
internal ballistics of a PCP airgun. Of these, the movement of the hammer is the only one
you can solve analytically between strikes, the rest must be solved numerically. For con-
venience, here is the full set.

2 
d xh kh 
= ------   apc +  dpc + L v – x h  x h  x v +  apc 
dt
2 mh 

2
d xh 
= 0 x h  x v +  apc 
2 
dt 
2 
d xv 1 
= – ------  F + k v   vpc + x v – L v   
dt
2 mv 

d p 1 dm dm p out 
= ------  p in – 
dt Vp  d t dt  

dT p 1 dm p in  dm p out  
-------- v 2in – -------- v 2out
1 1 
= ------ T – T + T – T + (EQ 43)
dt mp d t  in p 2c v  d t  out p 2c v  

d b 
1 dm b 
= ----------- – Ab vb b
dt Ab xb d t 


dT b 1 dm b  Ab 
T out – T breech + -------- v out – ------ p base v b – ---- d KE
1 2 1
= ------ 
dt m d t  2c  c c v dt
b v v 
2 
d xb F break 
= 0 ; p base  p nose + --------------
- 
dt
2 Ab 

2 
d xb Ab F break 
= ------  p base – p nose  – F ; p base  p nose + --------------
- 
2 M Ab
dt 

Internal Ballistics of PCP Airguns August 16, 2018 16


4.0 A case study
A case study is a practical way to explore the potential of a methodology when the dimen-
sionality of the problem is high, which is the case here. Table 1 lists the parameters of a
PCP airgun with a power of about 18 Joules1. Table 1 shows the parameters of a nominal
configuration - calculations will show variations around this nominal configuration.
TABLE 1. Nominal case parameters - initial temperature 288 K, initial pressure 1.01 bar.
Hammer Hammer Spring Valve Valve Spring Transfer Port Barrel and Pellet
Mass 30 g Free length 54 mm Mass 15 g Pre‐comp 20 mm Diameter 2 mm Caliber 4.5 mm
Attachment Free Numb coils 16 Head diam 8 mm Free length 25 mm Plenum vol 1 cm3 Pellet weight 0.546 g
Restitution 0.8 Wire diam 1.25 mm Stem diam 3 mm Spring const 19.61 N/cm Barrel length 300 mm
Outer diam 12.19 mm Seat modulus 3500 MPA Break force 19.61 N
Pre‐comp 10 mm Seat depth 4 mm Friction force 1.961 N
Cocking strk 24 mm Seat width 2 mm Seat depth 4 mm
Valve dist 54 mm Restitution  0.5
Young's mod 2E+05 GPA

Hammer strike and bounce. Let’s start by looking a what happens with the hammer and
the valve as the trigger is released. The next figure shows hammer and valve trajectories as
they intercept four times, with the fourth strike barely visible. The first is the strike that
fires the pellet, the remaining three are due to hammer bounce. Of the four strikes, only
three result in a successful opening of the valve. In the last strike, the hammer has insuffi-
cient energy to break open the valve. The pellet leaves after 7.96 ms since hammer release.
Notice that all the strikes after the first one happen after the pellet has left the barrel and do
not affect accuracy.

Although the bouncing of the hammer doesn’t affect accuracy (assuming the bouncing
happens after the pellet has left the barrel), it does affect the amount of air available in the
reservoir. For this reason, which will become clearer when we analyze the shot count
effectiveness of the gun, manufacturers struggle to eliminate hammer bounce through
mechanical arrangements that dissipate the hammer kinetic energy as it bounces back-
wards (known in the PCP airgun trade as “debouncing” devices.)

FIGURE 4. Hammer strike and two bounces for the


nominal case.
60
Valve stem
50

40

Position (mm) 30 Hammer
20

10
Pellet exit
0
0 5 10 15 20 25
Time since trigger release (ms)

1. In the UK and other European countries, ownership of airguns without a license is limited to 12 foot-
pounds of energy, or approximately 16 J.

Internal Ballistics of PCP Airguns August 16, 2018 17


Air mass discharge. The amount of air discharged at each strike and the number of times
the hammer bounces are influenced by the restitution coefficient of the hammer and valve
stem, and by the energy required to open the valve. With a restitution coefficient 0.8,
which is standard for steel on steel, there is less than 2% of wasted air beyond the first
strike. The next figure shows the mass ejected as a function of the strike sequence assum-
ing e = 1 , or 100% restitution.

FIGURE 5. Mass ejected from reservoir as function of


bounce sequence for 100% restitution.
300

250

200
Hammer restitution coeff = 1
Mass ejected
150
(mg)
100

50

0
1 2 3
Bounce sequence

Valve opening from the pellet perspective. To gain a better sense for the valve opening
and mass transfer during the first strike, let’s look at the position of the hammer and valve
stem as a function of the pellet displacement in the barrel. As is clear from Figure 6, most
of the valve movement (and the blast of high-pressure air) is over by the time the pellet
has moved a relatively short distance. This tells you that after the pellet has moved less
than 100 mm, the acceleration of the pellet will come from air that accumulated in the
transfer plenum and the barrel only. In a PCP airgun, high-pressure gas always flows in
one direction (from the reservoir into the transfer plenum and from the transfer plenum
into the barrel).

FIGURE 6. Hammer and valve stem position as


function of pellet position in barrel.
55

54.5
Valve stem
54

Position (mm) 53.5

53 Hammer
52.5

52
0 50 100 150 200 250 300 350
Pellet position in barrel (mm)

Internal Ballistics of PCP Airguns August 16, 2018 18


Valve lift and dwell in detail. Let’s now expand the first hammer-valve stem hit area in
Figure 4 to see the valve opening details - this is shown in Figure 7. Notice that the ham-
mer hits the valve stem three times during the time the valve remains open. As I men-
tioned earlier, this is one of the implications of elementary collision theory, where the
momentum exchange is instantaneous. An empirical verification of what Figure 7 shows
would be challenging, but highly desirable. However, even if a more refined collision
model changed the details in Figure 7, the overall features (dwell and lift), which depend
on hammer momentum and spring energy balance, would likely remain unchanged.

FIGURE 7. Valve opening when first hit by the hammer.


55
54.8
54.6
54.4
54.2
Valve stem
Position (mm) 54
Valve stem in 
53.8
closed position
53.6
53.4 Hammer
53.2
53
0 0.5 1 1.5
Time since hammer‐valve strike (ms)

Pellet velocity and reservoir pressure. It is well-known that the pellet velocity of an
unregulated PCP airgun is a concave function, which I will call the velocity curve, with a
maximum that defines the useful range of reservoir pressure. The shape of this curve is
typically attributed to the difference in pressure forces between the side of the valve
exposed to the reservoir, and the side exposed to the transfer plenum - if the reservoir pres-
sure is too high, the valve won’t open enough, if it is too low, the mass and enthalpy flows
wont be as large. This intuitive explanation is essentially correct, but there are complicat-
ing issues. Consider what happens when you change the pressure of the reservoir, and,
parametrically, also change the pre-compression of the hammer spring and the volume of
the transfer plenum. Figure 8 shows the velocity curve for three spring pre-compression
values. Notice that decreasing pre-compression shifts the maximum to the left. If you
define the operating rage by a drop of no more than 10 m/s, the operating pressure range,
in this particular case, would extend from approximately 140 bar to 220 bar.

Internal Ballistics of PCP Airguns August 16, 2018 19


FIGURE 8. Velocity curve for three values of adjustable pre-compression values (no preset pre-
compression) - operating range for 10m/s drop is between 220 and 140 bar.
300
251 m/s
250

200 10 mm pc
Velocity
150 5 mm  pc
(m/s)
100 No pc
Operating
50 range

0
0 50 100 150 200 250 300 350
Reservoir pressure (bar)

This tells you pre-compression has a significant effect, not just on maximum power, but
on the pressure where the maximum power obtains. However, as we will see later, results
also show that pre-compression as an independent parameter loses effectiveness as its
value increases.

Equally important is the transfer volume, as Figure 9 shows. The net effect of increasing
the transfer volume is to shift the velocity curve to the left, with a significant effect on the
maximum location.

FIGURE 9. Velocity curve for three transfer plenum


volumes (the nominal case is 1 cm3).
350

300 0.25 cm3
250

200 1 cm3
Velocity
(m/s) 150
4 cm3
100

50

0
0 50 100 150 200 250 300 350
Reervoir pressure (bar)

An alternative form of the velocity curve is the representation of projectile velocity as a


function of shot count, rather than reservoir pressure. To make the transformation from
total pressure to shot count, you need to specify the reservoir volume and the discharge
process. Unlike the firing cycle, whose time scale is of the order of milliseconds and can
be characterized as adiabatic, the time scale of the reservoir discharge if of the order of
hours; this suggests an isothermal process is adequate to represent the reservoir state dur-
ing discharge.

Internal Ballistics of PCP Airguns August 16, 2018 20


Assuming an isothermal discharge process, denote by m i the single-shot mass ejected
from the reservoir when the reservoir pressure, p 0 , is such that p 0 i + 1  p 0  p 0 i . The
reservoir pressure drop provoked by the release of m i is,

RT 0
p 0 = m i --------- , (EQ 44)
V

where V is the volume of the reservoir. The number of shots, n i , that cause the reservoir
pressure to drop from p 0 i + 1 to p 0 i , is, to first order in p 0 i – p 0 i + 1 ,

p 0 i – p 0 i + 1
n i = ------------------------------
-. (EQ 45)
p 0

The mapping of a function of reservoir total pressure, g , to one of shot count, f , is,

1 V
f  n i  = g  p 0 i  ----- ------- , (EQ 46)
m i RT

p 0
where p 0 i = p 0 i – p 0 i + 1 . For best accuracy, m i should be evaluated at p 0 i + ----------i
2
(this guarantees second-order accuracy.) To complete the transformation from total pres-
sure to shot count, you can also determine the gas expended in bounces beyond the first
strike1.

The plot in Figure 10 assumes the reservoir volume is 200 cm3, and that there is no bounc-
ing beyond the first strike.

FIGURE 10. Velocity curve as function of shot count, drop of 10


m/s or less, assuming no bounce after first strike - reservoir
volume = 200 cm3, pressure range 140 to 220 bar.
270

260

250

240
Muzzle vel.
(m/s) 230

220

210

200
0 20 40 60 80 100
Shot count

1. You can achieve the same result by simulating a large number of sequential shots.

Internal Ballistics of PCP Airguns August 16, 2018 21


Pressure and temperature. Figure 11 shows the time evolution of total pressure in the
transfer plenum and the average static pressure in the barrel. Keep in mind that due to the
Lagrange gradient approximation the breech pressure and the pressure right behind the
pellet are slightly different from the average pressure. Two notable features are that the
barrel pressure reaches a maximum before the plenum, and that about half of the plenum
pressure rise occurs under choked flow. Remember that under choked flow the geometri-
cal details of discharge are not relevant. In a PCP airgun, the effects of the nozzle geomet-
rical details are felt indirectly, through the valve opening dynamics (EQ 9).

Notice that in this particular configuration of parameters, the transfer port discharge
becomes supercritical when the pellet has spent about half of its residence in the barrel.
The reason why the pressure peak is reached earlier in the barrel is the relieving effect of
the pellet movement, as will become clear in the next section.

Notice that the maximum pressure in the transfer plenum comes close to the pressure in
the reservoir (175 bar in this case). Qualitatively, the pressure behind the pellet evolves in
a somewhat similar fashion as in a firearm [16].

FIGURE 11. Pressure history in plenum and barrel -


supercritical phase in barrel too narrow to be visible.
Supercritical 
200
reservoir discharge
180
160
140 Supercritical
120 breech discharge Transfer channel
Pressure (bar) 100
80
60
Barrel
40 Valve open
20
0
0 0.5 1 1.5 2
Time since hammer strike (ms)

The temperature evolution in the transfer plenum and in the barrel, shown in Figure 12, is
interesting because it affords you an opportunity to check a fundamental result of thermo-
dynamics. The first thing to notice is that, over a good part of the firing cycle, the temper-
ature in the transfer plenum and in the barrel is higher than in the reservoir. This may seem
surprising, but it is a well-known from Thermodynamics that when a large vessel dis-
charges adiabatically into an enclosed insulted vessel, the stagnation temperature rises by
a factor equal to the ratio of specific heats,  , which in the case of air is 1.4 [18].

The theoretical maximum of the temperature, both in the plenum and in the barrel, can
never exceed 1.4 times the stagnation temperature of the environment where the gas origi-
nates. The gas in the plenum originates in the reservoir, and the gas in the barrel originates
in the plenum. This means the temperature in the transfer plenum will reach, in general, a
higher value than in the reservoir, and in the barrel a higher value than in the plenum.
These two boundaries are beautifully illustrated in Figure 12.

Internal Ballistics of PCP Airguns August 16, 2018 22


If you are inclined to deeper thinking, you may want to ponder whether a concatenation of
discharge spaces leading up to the transfer port wouldn’t be a way to increase the gun
power and by how much.

FIGURE 12. Temperature evolution in transfer plenum and


barrel.
600 Max barrel adiabatic temp.

500 Barrel
Max TC adiabatic temp.
400
Transfer
Temperature
300 channel
(K)
200

100

0
0 0.5 1 1.5 2
Time since hammer strike (ms)

Pellet acceleration and position. In a PCP airgun the pellet movement starts with zero
acceleration - as shown in Figure 13. This is expected, since it takes some time to build up
enough breech pressure for the pellet to move.

FIGURE 13. Pellet velocity as function of time during transit in


barrel.
300
Valve open
250

200
Pellet
150
Velocity (m/s)
100

50

0
0 0.5 1 1.5 2
Time since hammer strike (ms)

Figure 14 shows the pellet velocity as a function of pellet travel within the barrel. Notice
that in this particular configuration of parameters, the valve remains open over a signifi-
cant fraction of the pellet residence time.

Internal Ballistics of PCP Airguns August 16, 2018 23


FIGURE 14. Pellet velocity as function of pellet travel in the
barrel - valve closes when pellet travels less than a third of
the barrel length.
300
Valve open
250

200
Pellet 
150
velocity (m/s)
100

50

0
0 50 100 150 200 250 300
Pellet position in barrel (mm)

Pre-compression and transfer volume. As Figure 15 illustrates, the intersticial volume


has a very significant impact on the pre-compression effectiveness in changing the pellet
exit velocity. To accommodate the range of pre-compression in Figure 16, our nominal
case was fitted with a 64 mm spring (same other dimensions and same number of coils.)
The case of 1.0 cm3 transfer volume shows what looks like a change of regime in the
effect of pre-compression, where beyond a certain value the pre-compression effect
appears to saturate. The reason for this saturation becomes clear when you take another
look at Figure 11. Once flow through the transfer port chokes, enhancing the reservoir dis-
charge will lose effectiveness in transferring mass and energy to the breech.

FIGURE 15. Compression effect on pellet velocity for three


transfer volumes.
300
0.25 cm3
250

200 1 cm3

Velocity
150
(m/s) 4 cm3

100

50

0
0 5 10 15 20
Pre‐compression (mm)

Velocity and barrel length. The pellet exit velocity grows monotonically with barrel
length until friction and nose work cause the velocity to drop. Figure 16 shows that the
intersticial volume causes a velocity drop by an almost uniform amount for any barrel
length.

Internal Ballistics of PCP Airguns August 16, 2018 24


FIGURE 16. Effect of barrel length for two transfer volumes.
350

300 1 cm3
250

200 4 cm3
Pellet vel.
(m/s) 150

100

50

0
150 250 350 450 550
Barrel length (mm)

Measures of efficiency. An efficiency measure quantifies how well the energy stored in
the reservoir is converted into kinetic energy at the muzzle. Two interesting aspects of the
efficiency of a PCP airgun are the effectiveness of the gun at extracting energy from a
blast of air, and the effectiveness of the gun in allocating a volume of air to the production
of kinetic energy. The former, which I will call thermal efficiency, is a true efficiency in
that it is dimensionless. The latter, which I will refer to as volumetric efficiency, is better
characterized as a figure of merit, since it has dimensions. There is an issue as to whether
or not a measure of efficiency should be corrected for friction losses, which are not the
“fault” of the gun. In the discussion here I made no such correction. You must take into
account two factors when interpreting efficiency numbers. One is the adiabatic assump-
tion of this model, the other is the number of bounces the hammer undergoes. In the
results that follow, I assume there is no hammer bounce after the first strike - this, plus the
absence of heat transfer means the efficiency numbers given here are optimistic. Correct-
ing for mass ejected during bouncing after the first strike is very simple, however.

Figure 17 shows how the thermal efficiency changes as a function of barrel length and
intersticial volume. The almost parallel position of the two curves is expected. Notice the
enormous impact of intersticial volume - when the transfer volume is small, these compu-
tations suggest that, in an adiabatic framework, the thermal efficiency of a PCP airgun can
be substantially higher than the efficiency of a spring-piston airgun [1].

FIGURE 17. Thermal efficiency versus barrel length for two


transfer volumes.
40
1 cm3

30

Thermal
4 cm3
Efficiency 20
(%)

10

0
150 250 350 450 550
Barrel length (mm)

Internal Ballistics of PCP Airguns August 16, 2018 25


The thermal efficiency tends to rise with reservoir pressure, as Figure 18 shows. The
changes are neither monotonic nor very significant.

FIGURE 18. Thermal efficiency versus reservoir pressure for


the nominal case with transfer volume 1 cm3.
40

30

Thermal
Efficiency 20
(%)

10

0
0 50 100 150 200 250 300 350
Reservoir Pressure (bar)

The volumetric efficiency, expressed in Joules per cubic centimeter of air at ambient tem-
perature and pressure, essentially conveys the same information as the thermal efficiency,
and you would expect the plots to be similar, as Figure 19 demonstrates. Which of these
two forms of efficiency to use is a matter of taste. The volumetric efficiency has the appeal
that you don’t need to think in terms of internal energy flow. The volumetric efficiency has
the (purely coincidental) advantage that, when expressed in American or Imperial units,
typical values are of order one foot-pound of energy per cubic inch, which provides a very
intuitive basis for comparison (in this case study, the volumetric efficiency is or order one
foot-pounds of energy per cubic inch.)

FIGURE 19. Volumetric efficiency versus reservoir pressure -


nominal case with transfer volume 1 cm3.

0.12

Volumetric 0.08
Efficiency
(J/cm3)
0.04

0
0 50 100 150 200 250 300 350
Reservoir Pressure (bar)

Impact on performance of the gas mass behind the pellet. As mentioned earlier, the
mass of the gas behind the pellet affects the gun performance in two ways, a) it changes
the temperature of the gas through the kinetic energy time rate of change, and b) it changes
the discharge environment of the transfer port and the base pressure of the pellet. All these
effects detract from the pellet muzzle velocity. For the sake of illustration, consider the
impact on pellet velocity as a function of the transfer port diameter - Figure 20 shows the

Internal Ballistics of PCP Airguns August 16, 2018 26


results. Notice that the velocity loss due to the onset of the Lagrange pressure gradient is
greater than the loss due to changes in thermal energy. For our nominal case, the total loss
of pellet velocity attributable to the gas inertia is about 10 m/s, a considerable number in a
competition environment. As you would expect, the importance of the gas kinetic energy
behind the pellet grows with the transfer port diameter.

FIGURE 20. Effect of gas inertia in barrel behind pellet -


nominal case with varying transfer port diameter.
16
Total loss
14
12 Loss due to
10 pressure
Pellet vel. gradient
8
loss (m/s)
Loss due to 
6
KE impact on
4 thermal energy
2
0
1 1.5 2 2.5 3
Transfer port diameter (mm)

Energy budget. Table 2 shows the energy allocation at pellet exit time for the nominal
case. Notice that the thermal energy remaining in the barrel by the time the pellet leaves is
larger than the kinetic energy of the pellet.
TABLE 2. Energy budget and allocation at pellet exit time,
nominal case.
Fraction 
Energy at time of pellet exit Joules
(%)
Thermal energy in transfer plenum 18.1 26.8
Thermal energy in the barrel 27.1 40.1
Gas kinetic energy in the barrel 2.6 3.8
Energy dissipated by pellet nose 0.5 0.7
Pellet kinetic energy 18.7 27.6
Friction energy dissipatd by pellet 0.6 0.9
Total energy spent 67.5
Energy released from reservoir 67.5

5.0 Concluding remarks


This study shows that an inviscid quasi-stationary framework captures the most salient
aspects of a PCP airgun, such as optimal reservoir pressure, effect of hammer spring stiff-
ness, impact of gas inertia behind the projectile, and interstitial volume. A number of
issues remain open to further research and experimental verification, most notably the
modeling of the hammer-valve interaction and the role of kinetic energy in the transfer
channel. Given the high dimensionality of the problem, this study was focused on varia-
tions around a nominal configuration. There is ample potential for enriching the methodol-
ogy with additional features, such as accounting for viscosity and non-stationary gas
dynamics.

Internal Ballistics of PCP Airguns August 16, 2018 27


6.0 References
[1] D. Tavella, “Internal ballistics of spring-piston airguns”, April 2015 (downloadable
from Researchgate)

[2] D. Tavella, “Spring buzz and failure in spring-piston airguns”, June 2015 (download-
able from Researchgate)

[3] D. Tavella, “Dieseling in spring-piston airguns - a conceptual analysis”, Feb 2015


(downloadable from Researchgate)

[4] http://www.fortwiki.com/Battery_Dynamite_(3)

[5] M.C. Ortner, “The Austro-Hungarian Artillery from 1867-1918: Technology, Organi-
zation, and Tactics”, Verlag Militaria, 2007

[6] https://ares.jsc.nasa.gov/orbital_debris/hvit/impact/light-gas-guns.html

[7] M. Denny, “Internal ballistics of an air gun”, The Physics Teacher, (40), Feb. 2011

[8] C.E. Mungan, “Internal ballistics of a pneumatic potato gun”, Eur. J. Physics, (30),
March 20092

[9] J.Barlow, S. Pond, and M.J. Madsen, “Initial model of a pneumatic air cannon”, WJP
(2), Dec. 2009

[10] H. Horak et al, “Prediction of the air gun performance”, Advances in Military Tech-
nology, (9), 1, June 2014

[11] F. Plasard et al, “Analysis of a single-stage compressed gas launcher behavior - from
breech opening to sabot separation”, on-line report.

[12] B.A. Hardage and F.C. Todd, “Design and construction of a helium gas gun for
hypervelocity impact”, http://digital.library.okstate.edu/OAS/oas_pdf/v45/p129_138.pdf

[13] H. Hertz, “Ueber die Beruehrung fester elastischer Koerper”, J. Reine Angewandte
Mathematik, 92, 1881, 155

[14] W. Goldsmith, “Impact - The Theory and Physical Behavior of Colliding Solids”,
Dover Publications, 2001

[15] D.A. Jobson, “On the flow of a compressible fluid through orifices”, Proceedings of
the Institution of Mechanical Engineers, June 1955

[16] J. Corner, “Theory of the interior ballistics of guns”, John Wiley and Sons, 1955

[17] M. Acosta and J. Tamagno, “Modelado cuasi-unidimensional del movimiento de un


piston en un tubo”, XIV Congreso sobre metodos numericos y sus aplicaciones, Argen-
tina, 2004

Internal Ballistics of PCP Airguns August 16, 2018 28


[18] W.C. Reynolds and H.C. Perkings, “Engineeering Thermodynamics”, McGraw-Hill,
Inc., 1977

Internal Ballistics of PCP Airguns August 16, 2018 29

Das könnte Ihnen auch gefallen