Download as pdf or txt
Download as pdf or txt
You are on page 1of 111

Hybrid Rocket Propulsion System

Senior Design II: Mech 4035-002


Professor: Doug Gallagher
05/05/2023

Project Lead: Alejandro Miramontes


Team Leads: Sean Zygmunt, Willis Begaye
Senior Team Members: Garrett Price, Luis Ortiz, Ryan Boulden, Erin Roth, Michael Gisch.
Joel Luna, Marco Diosdado
Abstract (Alex) - Edit Passed tense now

The RocketLynx team was first established to produce a complete, functional

student-researched, and designed rocket system meeting the requirements for competing

in the Spaceport America (SA) cup. This year’s team, RocketLynx V, was focused on

improving the hybrid propulsion system so that future teams can begin airframe

integration, working towards the final goal of entering a vehicle into the SA cup.

The current iteration of the team was fulfilling the ultimate goal of entering a vehicle into

the SA cup by battling the goals laid out for this year. The team planned to design, analyze,

manufacture, and test a new motor prototype capable of reaching a thrust of 250 [lbf.]

while achieving a 5:1 thrust-to-weight ratio. This new engine iteration continued to utilize

nitrous oxide (N2O) as the oxidizer and acrylonitrile butadiene styrene (ABS) as the solid

fuel. Redesigns took place for all subsystems to meet design requirements but also to begin

moving away from a test configuration into an engine that can be implemented into an

airframe. Specifically, we redesigned the combustion chamber, pressure vessel, nozzle,

ignition system, and looked to improve the avionics system along with the fuel grain.

Once the engine was designed and manufactured, the team used reliable data acquisition

(DAQ) to collect data from a static hot-fire test. With improvements to the ignition system,

data was collected more efficiently as the team will be able to ignite the system multiple

times. Along with this, modifications for the avionics circuitry allowed for reliable

connections between each electronic component but also reliable connections to the DAQ

system.
Table of Contents (Alex, Willis, Sean)

Abstract x

1. Introduction x
1.1 Hybrid Propulsion Model x
1.2 Project Goals x
1.3 Current Stat of Project x

2. Design Adaptation and Analysis x

2.1 Propulsion Team x


2.1.1 Fuel Grain Redesign x
2.1.2 Ignition Port Redesign x
2.1.3 Combustion Chamber Redesign x
2.1.4 Nozzle Flange Redesign x
2.1.5 Nozzle Current State x
2.2 Oxidizer Team x
2.2.1. Pressure Vessel Current State x
2.2.2 Injector Current State x
2.3 Avionics Team x
2.3.1 Master On/Off Redesign x
2.4 Test Stand x

3. Manufacturing Methods x
3.1. Ignition Circuit x
3.1.1. Fuel Grain Manufacturing x
3.1.2 Combustion Chamber Manufacturing x
3.1.3 Nozzle Manufacturing x
3.1.4 Nozzle Flange x
3.2 Oxidizer Team x
3.2.1 Pressure Vessel Manufacturing x
3.2.2 Injector Manufacturing x
3.2.3 Injector End Cap Manufacturing x
3.2.4 Oxidizer Cap Manufacturing x

4. Test Stand x

5. Avionics System x

5.1 Ignitor Circuit x


5.2 Flow Metering and Control x

5.3 Load Cell Calibration x

6. Test Preparations x

7. Conclusion x

8. Budget x

9.Gantt Chart x

References x

Appendices x

List of Figures (Alex, Willis, Sean)


    Figure 2.1.1.1: Previous Model of the Fuel Grain x
    Figure 2.1.1.2:Final design of the fuel grain x
Figure 2.1.1.3: Fuel grain glued with ABS cement x
Figure 2.1.1.4: Metal Pinch Clamps to secure the liner to the fuel grain. x
Figure 2.1.1.5: Fuel grain nozzle side (first half)) x
Figure 2.1.1.6: Fuel grain nozzle side (second half) x
Figure 2.1.1.7: Fuel grain injector side x
Figure 2.1.1.8: CFD analysis (Pressure Cut Plots) x
Figure 2.1.1.9: CFD analysis (Velocity Flow Trajectories) x
Figure 2.1.2.1: Previous Ignition 3-Port Design x
Figure 2.1.2.2: Potassium Nitrate used in Rocket Candy Mixture. x
    Figure 2.1.2.3. 3-Port Design filled with 2:1 Rocket Candy Mixture. X
Figure 2.1.2.4. Ignition Cap 2-Port Design. X

Figure 2.1.2.5. 2-Port Design with Black powder top layer. X

Figure 2.1.3.1: Combustion Chamber Model x

Figure 2.1.3.2: Finite Element Analysis of Combustion Chamber. X

Figure 2.1.3.3: Main Combustion Chamber Assembly. X


Figure 2.1.3.4: Computational Fluid Dynamics Analysis of Combustion Chamber. X

Figure 2.1.3.5: Silica Heat Insulating Liner x

Figure 2.4.1: Test Stand FEA Von-Mises Stress Plot x


Figure 3.1.1.1: Original Prusa MK3 3D Printer x

Figure 3.1.3.1: Cam Simulation of Outside Profile of Nozzle x

Figure 3.1.3.2: Facing Operations of Nozzle x

Figure 3.1.3.3: Boring of Nozzle Convergent Section x

Figure 3.1.3.4: Boring of Nozzle Divergent Section x

Figure 3.1.3.5: O-Ring Glands Operation x

Figure 3.2.1.1: OD Grooving Operation x


Figure 3.2.1.2: ID Drilling Operation x
Figure 3.2.1.3: ID Boring Operation x
Figure 3.2.1.4: Pressure Vessel Body x
Figure 3.2.2.1: Grooving Operation of Injector Face x
Figure 3.2.2.2: Boring Operation of Injector Side (Top) x
Figure 3.2.2.3: Boring Operation Combustion Chamber Side (Bottom) x
Figure 3.2.2.4: Profile Roughing Oxidizer Cap (Tapered Hole Side) x
Figure 5.1.1: Ignition Circuit Housing Unit x
Figure 5.2.1: PID Tuner Function x

List of Tables (Alex, Willis, Sean)

Table 2.1.5.1: Nozzle Dimensions x


1. Introduction (Alex)

1.1 Hybrid Propulsion Model (Alex) - Edit to Past tense

Hybrid rocket propulsion systems are designed following the bipropellant concept, where

the most common configurations have an oxidizer source stored in its liquid phase and a

fuel source stored in its solid phase. Thrust is produced when the solid fuel source goes

through a combustion process, which is enhanced by the oxidizer flow through the

combustion chamber. Before the exhaust gas exits the combustion chamber, a transition

from subsonic to supersonic occurs with the nozzle's help at the chamber's exit. The nozzle

is specifically designed to maximize the thrust output potential of the engine. In general,

avionics systems are implemented to control the ignition, oxidizer flow and receive engine

data. Research into hybrid rocket propulsion systems has been done since the 1930s and

has continued due to the system's unique advantages. Some advantages are start-stop and

restart capabilities, an ability to throttle the engine during operation, and a more

comprehensive range of cost-effective fuel options. Previous RocketLynx teams have done

extensive research within hybrid rocket propulsion engines while showing promising

results. Gilmour Space Technologies, based in Australia, has developed a hybrid rocket

engine that can reach 90 [kN] of thrust within testing. As research for hybrid rocket

propulsion systems becomes readily available, innovations can begin to improve the model

resulting in a promising future within rocketry and aerospace technologies.


1.2 Project Goals (Alex) - Edit to Past Tense

As the fifth iteration of the RocketLynx team, RocketLynx V innovated a new iteration of the

rocket engine prototype to begin preparing future teams for airframe implementation. The

team followed strict goals and requirements within each subsystem to achieve this. The

team was split into three sub-teams: the propulsion, the oxidizer, and the avionics teams.

Each sub-team focused on specific aspects of the system, while integration will be handled

as a team. The project lead and the team leads ensured requirements were set, and team

members worked to develop their components to meet set requirements.

The first goal that the RocketLynx V team faced when designing the new rocket prototype

was achieving a peak thrust of 250 [lbf.] during static hot-fire testing. The team wanted to

innovate the fuel grain and nozzle to achieve this desired thrust. The next goal the team

tackled was achieving a thrust-to-weight ratio of 5:1. While innovations to the fuel grain

and nozzle improved thrust output, new designs for the combustion chamber and pressure

vessel were implemented to minimize the overall weight. Finally, the team worked towards

implementing a reliable ignition source allowing the team for on-command ignition at any

stage. As the ignition source requires safety features, the avionics team will work towards

improvements within the avionics system to allow for secure circuitry. Not only will this

allow for reliable data acquisition, but it will also allow setup to be more efficient

throughout testing procedures. With innovations for all subsystems, the team hopes to

produce a system that can be taken by future RocketLynx teams to begin airframe

implementation.
1.3. Current State of Project (Alex) - Edit to end of year

With the project having all core designs, analysis, and manufacturing processes, the team

has transitioned into the manufacturing and assembly phase. New designs and analyses

have been adapted to the project based on assembly requirements and manufacturing

considerations to improve the functionality of each component. All new designs and

analyses are discussed within this paper and serve as the final design iteration of each

component that has undergone a design review. Any component not mentioned as having a

design update has already gone through the final iteration and was ready for manufacturing

at the beginning of the Spring 2023 semester.

Currently, manufacturing has been completed on all components. It includes the order of

operations for manufacturing, purchasing of all stock materials, and the development of

manual and computer-aided processes for machining. The team has decided to work with

similar machining operations for all components to ensure ease of manufacturing. With

this, the team has collaborated and shared information for machining and best practices for

maximum efficiency. These processes are discussed within this paper while also discussing

the final results for components.

Within the DAQ/avionics side, initial calibration of all sensors was completed and finalized.

The team is actively worked to adapt all sensors and data acquisition tools to meet new

requirements in the system. The sensors being used have not changed from previous

iterations and include pressure transducers, a load cell, and a thermocouple. Once

calibration was completed, the team began moving on to the testing and assembly of all

data acquisition instruments.


As the team finalized manufacturing, assembly, and calibration, the team was ready to begin

testing the full-scale hot-fire static test. All test procedures are discussed in length within

the sections below. These discussions give an initial idea for verifying components'

functionality while ensuring team safety during all testing operations.

2. Design Adaptation and Analysis (Alex)

Section 2 of this report discusses all component redesigns since the beginning of the Spring

2023 semester. The reasons for changes and resulting analysis are discussed to show each

component's functionality. For any component that has not gone through any changes, this

section discusses the current state of the component and shows the resulting analysis used

to achieve the conclusions.

2.1 Propulsion Team

2.1.1. Fuel Grain Redesign (Marco)

The fuel grain was considered one of the most vital aspects of the combustion chamber

because this was the material combusted in conjunction with the nitrous oxide (NO2). The

team reviewed previous fuel iterations and determined that the helical port produced a

more homogeneous thrust output. Initially, we considered a star shape port for the fuel

grain, but that configuration exceeded the (O/F) ratio. Furthermore, we analyzed different

materials to manufacture the fuel grain, such as paraffin wax, HTPB, and ABS. However, we

decided that ABS was the optimal material because it was easy to manufacture and had

similar properties to HTPB.


Furthermore, reviewing material properties, such as tensile strength (σu), ignition

temperature, cost, and ease of access, helped determine that ABS was the most appropriate

material. The tensile strength of all materials was similar, yet ABS yielded a higher tensile

strength combined with easier accessibility via additive manufacturing. Therefore, we

reconfirmed ABS as the most suitable material for this rocket design.

The team conducted calculations and created CAD models, and our results indicated that

the surface burn area (Aburn) would only slightly decrease with minimal effect on the total

thrust output.

The helical port design was configured to ensure an optimal oxidizer to fuel ratio (O/F)

based upon calculations of Aburn, the radius of the port (rport), fuel regression rate (ṙ), length

of fuel grain (LFuel), the cross-sectional area of the port (Aport), and mass flux of the oxidizer

(Gox). The Aburn of the fuel grain is calculated from equation (1) and is a function of time

since the fuel depletes as combustion is prolonged. The variables used can be described as

follows:

Aburn = changing area over time

rport = radius of the port

LFuel = length of the port

c = circumference of the helical port

2
𝐴𝑏𝑢𝑟𝑛 = 2π × 𝑟𝑝𝑜𝑟𝑡 × 𝐿𝑓𝑢𝑒𝑙 × π ×𝑐 (1)
The fuel regression rate was a factor in determining the thrust output and Aburn as time

progressed. This constant relied upon Gox and was calculated from equation (2) Where:

Gox = mass flux of the oxidizer

ṁtot = total mass flow rate

Aport = area of the helical port

𝑚̇𝑡𝑜𝑡
𝐺𝑜𝑥 = 𝐴𝑝𝑜𝑟𝑡
(2)

The fuel grain had to be redesigned to an overall length of 16.072 [in], and the outer

diameter was modified to 3.25 [in] The inner diameter of the helical port, 0.59 [in],

remained unchanged. The previous length of the fuel grain was 13.45 [in], the outer

diameter was 3.45 [in], and there were no grooves present in the model (Figure 2.1.1.1).

Figure 2.1.1.1: Previous model of the fuel grain


However, the nozzle suffered some changes in its length, spacers were added later in the

process, and the combustion chamber inner diameter changed slightly; consequently, the

fuel grain’s geometry had to be modified (Figure 2.1.1.2).

Figure 2.1.1.2: Final design of the fuel grain

A simple interlocking key system was designed to assemble all the parts, including the

igniter cap. ABS cement was used to prevent the parts from moving once all the

components were inserted in the combustion chamber (Figure 2.1.1.3).


Figure 2.1.1.3: Fuel grain glued with ABS cement

A carbon-fiber sleeve covered the fuel grain, and a groove was designed to insert a metal

clamp and prevent the sleeve from sliding (Figure 2.1.1.4). The width and depth of the

groove were 0.43 [in] and 0.30 [in], respectively. The engineering drawings confirm the

changes were made in the fuel grain (Figure 2.1.1.5, Figure 2.1.1.6, Figure 2.1.1.7)
Figure 2.1.1.4: Metal Pinch Clamps to secure the liner to the fuel grain.

Figure 2.1.1.5: Fuel grain nozzle side (first half)


Figure 2.1.1.6: Fuel grain nozzle side (second half)

Figure 2.1.1.7: Fuel grain injector side

After testing the hybrid rocket propulsion system, there were no remains of the fuel grain.

Our assumptions for this outcome were that the fuel grain was manufactured with a 70%

infill rather than 90% or 100%. Moreover, the nitrous oxide (NO2) mass flow rate was

calculated to be 1.0 [kg/s], and the pressure vessel was only filled 50 % of its capacity
because of conditions beyond our control, which reduced the mass flow rate to 0.3 [kg/s].

As a result, the combustion process lasted 24 seconds instead of 8 seconds, and the fuel

grain disintegrated. Our CFD analysis of the fuel showed the predicted results for a mass

flow rate (ṁ) of 1.0 [kg/s] (Figure 2.1.1.8 and Figure 2.1.1.9).

Figure 2.1.1.8: CFD analysis (Pressure Cut Plots)

Figure 2.1.1.9: CFD analysis (Velocity Flow Trajectories)

2.1.2. Ignition Port Redesign (Willis)

The ignitor port was introduced to achieve reliable ABS fuel grain material ignition.

Limiting the extension of the combustion chamber and overall project cost was a priority.

Therefore, we decided to remove the fuel grain's first 1.25 [in] to create the ignitor port.
Previous design iterations included three grooves, an internal ring, and ports for the

ignition wires (Figure 2.1.2.1).

Figure 2.1.2.1. Previous Ignition 3-Port Design.

The ports and grooves of the 3-Port Design were filled with a Rocket Candy mixture. This

mixture was created by combining Potassium Nitrate (Figure 2.1.2.2) and ordinary table

sugar in a 2:1 ratio dissolved in hot water.


Figure 2.1.2.2. Potassium Nitrate used in Rocket Candy Mixture.

Once the potassium nitrate and sugar were dissolved, medium to low heat was applied with

constant stirring until a thick paste had formed and most water evaporated. This mixture

was added to the 3-Port design (Figure 2.1.2.3)

Figure 2.1.2.3. 3-Port Design filled with 2:1 Rocket Candy Mixture.
Testing of the 3-Port Design revealed opportunities for improvement in the design.

Problems arose in printing the inside “groove” ring requiring support material that could

not be adequately removed. The remaining material impeded a complete filling with the

Rocket Candy mixture and consequently did not burn around the ring to the other deposits.

The three ports intended for the ignition cap were not the correct diameter and did not

allow enough contact with the Rocket Candy mixture. The design of a new ignition port was

completed with previous testing results in mind (Figure 2.1.2.4).

Figure 2.1.2.4. Ignition Cap 2-Port Design.

This new design shows multiple changes: reduced to two ports, increased deposit area,

elimination of groove for sidewalls, integrated ignitor cap, and an outer groove ring. The

Avionics team design required a reduction from three to two ports, allowing only two

ignition wires to be utilized. Consequently, reducing ports required an increase in the

deposit area to allow the same amount of Rocket Candy. There are still ports for the ignitor

wire, but they are further integrated into the deposit area, allowing for increased contact

with the Rocket Candy. The sidewalls were introduced inside the design, so no support
material would be required in printing. Simplification improved the design and allowed a

groove between ports for even ignition. The outer groove ring was implemented to allow

the metal pinch clamp to secure the fiberglass sleeve. This ring required support material,

but since it was external, removal of support material was easier.

The “Rocket-Candy” mixture was adjusted to provide a more reliable ignition at this stage.

We were still using a 2:1 ratio of Potassium Nitrate: Sugar; however, we only laid an initial

layer of this mixture first. Black powder (2 [g]) was added to the remaining mixture and

placed as a top layer of the ignitor cap (Figure 2.1.2.4). Through various testing, we

discovered that this mixture ignites reliably once allowed to cure for 24 [hrs.]

Figure 2.1.2.5. 2-Port Design with Black powder top layer.

2.1.3. Combustion Chamber Current State (Erin/Michael)

The combustion chamber housing underwent optimization, given the internal dimensions needed

to fit the fuel grain tightly, heat-insulating liner, and nozzle. When designing the combustion

chamber, the driving factor was the ability to withstand the temperature and pressure of the

burning oxidizer and fuel grain. Aluminum 6061-T6 has a melting temperature of around 855 [K]

and becomes too weak to withstand our predicted internal pressures at around 550 [K]. Therefore,

the safety cutoff temperature was determined to be 825 [K], and a heat-insulating liner was needed.
The heat resistance of the fiberglass liner, aluminum 6061-T6, was able to withstand the

combustion temperatures we encountered during testing. The max internal pressure expected was

calculated to be 997.7195 [psi] using equation (3).

(𝐶*)*ṁ
𝑃𝑚𝑎𝑥 = 𝐴𝑡
(3)

With a given , the hoop stress () of 7366.4074 [psi.] was calculated using the dimensions from Table

2.1.3.1 and equations (4) and (5).

𝑟𝑜*𝑃𝑚𝑎𝑥
σ𝑙 = 2𝑡𝑤
(4)

σℎ = 2 * σ𝑙 (5)

Aluminum 6061-T6's ultimate tensile strength is 400030 [psi], so the material was

expected to withstand the internal pressures produced. The sizing of the combustion

chamber relied on a list of requirements that needed to be met. The design had to enclose

multiple elements; the liner, the pre and post-combustion spacers, the ignitor cap, the

nozzle, and the fuel grain to create optimal engine efficiency and adequate thrust output to

meet our goal of 250 [lbf]. The combustion chamber size also had to consider the injector

and nozzle cap. In order to reach the necessary performance and efficiency of the

combustion chamber, the sizing for all of the component's dimensions was calculated, and

these values are seen in Table 2.1.3.1. If the diameter for the combustion chamber is too

small or too big, it will not correctly fit the phenolic liner, fuel grain, or nozzle; all of these

components have been designed and calculated with our thrust output goal in mind.

Another integral part of combustion chamber sizing is the pre-combustion and

post-combustion chambers. The pre-combustion chamber must have ample room for the
oxidizer to complete atomization and ignition. If the space is too large, the momentum from

the liquid oxidizer will be lost, and the energy from the burning oxidizer will be lost. If the

space is not large enough, there would not be enough room for combustion. The

post-combustion chamber sizing is critical to allow proper boundary layer growth along the

fuel grain and complete the burn of the oxidizer/fuel grain mixture before entering the

nozzle. The optimal length-to-diameter ratio (L/D) has been determined using standards

from AIAA-2010-183 (Dennis et al., 2010). For the pre-combustion chamber, the L/D is 0.5,

and for the post-combustion chamber, L/D is 0.75. The values determined using these ratios

were 3.26 [in.] and 4.90 [in.], respectively. These dimensions can also be found in Table

2.1.3.1.

Table (2.1.3.1): Combustion Chamber Dimensions.

Dimensions Meters (m) Inches (in)

Chamber

Lcc = 0.6477 25.5000

OD, cc = 0.0991 3.9000

ro , cc = 0.0413 1.9500

tw , cc = 0.0051 0.2000

Pre-Combustion Chamber
Lprec = 0.0828 3.26

ID, prec = 0.067056 2.64

Post-Combustion Chamber

Lpostc = 0.1245 4.90

ID, postc = 0.067056 2.64

Flanges

tflange = 0.0305 1.2000

OD, flange = 0.1905 7.5000

With the sizing specifications from Table 2.1.3.1 in mind, a model of our combustion

chamber was created in SolidWorks, as seen in Figure 2.1.3.1. This model also included

flanges that securely attached the injector and nozzle to the combustion chamber. These

flanges did not add any additional length to the chamber because they were designed to

retain our length of 25.5 [in].


Figure 2.1.3.1: Combustion Chamber Model

In order to secure the injector cap and nozzle to the combustion chamber, we implemented

flanges as part of this year's design. Last year, RocketLynx IV used all threads to secure the

components of the rocket together. This year's flange design removed the need for the

all-threads and lessened the overall weight added for the securing faculties. The flanges and

the combustion chamber were manufactured as one piece in order to avoid the need for

welding. Welding lowers the safety factor of the system and, in the case of rockets, could

lead to a catastrophic failure especially being part of the combustion chamber. For this

reason, we determined that machining the combustion chamber and flanges from one solid

piece of aluminum 6061-T6 would be the safest method. Ensuring the flanges had a sealed

fit on either side of the chamber, O-rings were incorporated to guarantee that all pieces

were correctly mated together, establishing safety and preventing any loss of thrust output.

We used ASME standards for pipes and flanges to design the flanges (ASME B16.5, 2017) to

determine a class 600 rating flange. Using our calculated combustion chamber internal

diameter of 3.5 [in] as specified in table (2.2.1), we determined the remaining flange and

bolt dimensions necessary. Utilizing the ASME standards dimensions in Table 2.1.3.2, we

determined that the diameter of the flanges was 7.5 [in.], the bolt circles were 5.88 [in.],
and the thickness of the flanges was 1.2 [in.]. Incorporating a thickness larger than the one

specified by the ASME standards increased our safety factor and minimally affected the

weight of the combustion chamber. Due to the outer diameter of the flange being the widest

part of the combustion chamber, this dimension also determined the diameter of the

aluminum 6061-T6 rod we purchased to be machined.

Table (2.1.3.2): Combustion Chamber Flange Dimensions.

Flange Dimensions per ASME Standards (ASME B16.5, 2017)

Outer Diameter = 7.5 [in.]

Thickness = 1.2 [in.]

Bolt Circle = 5.88 [in.]

No. of Bolts = 8

Bolt Hole Diameter = 0.50 [in.]

A time-based analysis was utilized to calculate the pressure produced by the fuel grain for

up to a fourteen-second burn time. Using equation (3), the maximum pressure was

calculated to be 997.7195 [psi.] (6.8790 [MPa.]). This maximum pressure was then utilized

to calculate the following stresses: longitudinal, hoop, and von mises. Table (2.1.3.3) shows

the maximum stresses exerted through the combustion chamber for each stress both in

[MPa.] and [psi.].


Table (2.1.3.3): Maximum Pressures from Time-Based Stress Analysis

Stress Pressure [Mpa] Pressure [psi]

σl 33.535 4863.88

σh 67.070 9727.765

σv 58.084 8424.491

Of the three stresses that were calculated, the hoop stress exerted the most significant

pressure on the combustion chamber. The hoop stress (σh) and the ultimate strength of

Aluminum 6061-T6 (σu = 276 MPa) were utilized with equation (6) to compute the safety

factor of the combustion chamber. The result produced a safety factor of 4.11.

To confirm the time-based analysis results, the team used Finite Element Analysis (FEA) to

determine if the time-based analysis was accurate.

The Finite Elements Analysis aimed to determine if the material chosen for the combustion

chamber would allow yielding at any given point. The combustion chamber model was put

into a SolidWorks simulation, and pressures were applied based on the real-world scenario

it would experience during testing. The analysis determined the max contact pressure and

the max Von Mises stress of the combustion chamber to determine if there would be yield

during this real-world testing scenario. The combustion chamber consists of three parts:

the main body of the combustion chamber, the injector flange, and the nozzle flange. All

parts were manufactured using aluminum 6061-T6. However, the FEA model looks at the
main body of the combustion chamber only because the injector flange needs minor

adjustments.

Using the max pressure produced by equation (3), as well as the ambient pressure of

Watkins Colorado (the original expected testing site) on a typical warm day (Pamb = 12.149

[psi.]), the pressures were applied to the combustion chamber. The maximum pressure the

combustion chamber would experience was applied to the internal surface of the chamber.

In contrast, ambient pressure was applied to the external surface of the chamber. The

maximum pressure simulates the expected pressure in the combustion chamber due to the

combustion process being produced by the combustion process of the fuel grain with the

liquid oxidizer (N2O). The ambient pressure simulated the environmental pressure due to

the location of testing.

Since the test was a static fire test, boundary conditions were applied to meet the testing

requirements in the FEA simulation. The flange that will hold the injector plate and injector

was treated as a bounded condition, given that bolts and the run system will hold the

surface in place. Along with that, the body of the combustion chamber will have a clamp

holding the chamber to the test stand, so a bounded condition was added to the body to

mimic a clamp on the body. These two conditions held the combustion chamber fixed

during analysis as the stresses were calculated. Finally, a mesh was created on the model to

provide results. The mesh was set up at the highest quality to allow for the most accurate

results in the analysis.

With the analysis being based on a calculated and expected ambient pressure, it was

determined that the maximum Von Mises stress the chamber would experience is 70.946

[MPa.]. The only location where this maximum stress was seen was where the boundary
condition was placed, mimicking the clamp on the test stand. When looking closely, it is

seen that the internal combustion chamber was only experiencing a stress of 56.772 [MPa.]

- 63.859 [MPa.], as can be seen in Figure 2.3.1.

Figure 2.1.3.2: Finite Element Analysis of Combustion Chamber.

The primary combustion chamber assembly, represented in Figure 2.1.3.3, consists of the

injector flange, injector, combustion chamber, heat-insulating liner, fuel grain, nozzle, and

the pre/post-combustion chambers spacers.

Figure 2.1.3.3: Main Combustion Chamber Assembly.


The assembly was used to run a Computational Fluid Dynamics (CFD) analysis to determine

the flow of the liquid oxidizer through the combustion chamber. When setting up the CFD

analysis, choosing parameters to determine the fluid flow was essential. The first step to do

this was setting up the computational domain of the analysis. The computational domain

was set to start at the entrance of the flange holding the injector and end at the nozzle exit.

The computational domain is vital as that is where the fluid flow calculations will be

simulated. From there, the fluid was selected for analysis. Since SolidWorks does not have

Nitrous Oxide (N2O) as a fluid, Carbon Dioxide was selected as the gas going through the

system due to its similar properties to Nitrous Oxide. Finally, the boundary conditions

included in the analysis were the feed system's mass flow rate and Watkins, Colorado's

ambient pressure. The mass flow rate was set at 1 [kg/s] at the entrance of the combustion

chamber, while the ambient press by ure was set to 83765.07 [Pa.] at the exit. The results of

the CFD analysis showed that fluid flow through the combustion chamber was consistent

with expectations. While vortices are created throughout the combustion chamber, this was

expected due to the components sitting inside the chamber and having walls where the

fluid will encounter. The results of the analysis can be seen in Figure 2.1.3.4.
Figure 2.1.3.4: Computational Fluid Dynamics Analysis of Combustion Chamber.

In the previous iteration of RocketLynx, a phenolic liner was used as a heat-insulating layer

between the fuel grain and the aluminum shell of the combustion chamber. A

heat-insulating layer is crucial to designing a combustion chamber with our expected

temperatures. As the temperature of aluminum rises, the material becomes weaker.

Keeping as much heat inside the fuel cell also benefits the burn process. This year the team

increased the diameter of the combustion chamber. However, due to the lack of availability

in size and budget needed to accommodate this change, a liner made of phenolic was not

viable.

After extensive research, a handful of potential heat insulators, namely Fiberglass or Silica,

have been found. This material had higher thermal conductivity, as seen in Table (2.1.3.4),

so the new thickness needed for the insulating liner was also adjusted. We aimed to ensure

the aluminum shell did not reach over 550 [K]. At this point, aluminum’s yield strength is

cut in half. While the maximum stresses on the aluminum are still well below this value,

avoiding structural failures is imperative. Safety is always the most critical factor.

Table (2.1.3.4): Thermal Conductivity of Insulating Material

Thermal Conductivity [W/m-K]

Phenolic = 0.25
Silica = 1.3

Another variable was whether to buy a premade expandable heat-insulating sleeve or use

an X-Winder to create our own. The benefit of the X-Winder is the ability to lay the

insulation directly on the fuel grain. However, using this method also requires resin to

solidify the wrap. Resins that can withstand our maximum temperature of 825 [K] are not

easy to come by and expensive when they are found. With the expandable sleeves, a

potential issue is a sleeve slipping off when inserting the fuel grain into the combustion

chamber. The sleeve was adequately secured; however, finding a diameter of 3.5 [in.] was

also challenging. A 5-foot sample was sent to us of the below material in Figure 2.1.3.5.

Figure 2.1.3.5: Silica Heat Insulating Liner


2.1.4. Nozzle Flange Current State (Erin)

Figure 2.1.3.3: Final Design for the Nozzle Flange Cap

The nozzle flange was designed to have a thickness of 1.2 [in] to match the thickness of the

other flange components. The design also contained the same bolt circle, bolt number, and

bolt numbers located on table 2.1.3.2. Although analysis was done on having an angular exit

hole to accompany the angle of the nozzle the best results came from having an exit hole

that had a consistent inner diameter of 2.4 [in].

2.1.5. Nozzle Summary/Current State (Joel)

The supersonic nozzle for a hybrid rocket propulsion system is responsible for achieving a

high degree of conversion of enthalpy to kinetic energy created by an exothermic reaction

between the solid fuel and oxidizer from the combustion chamber. The ratio between the

inlet and exit pressures in all rocket propulsion systems will be large enough to induce

supersonic flow.
Before designing a nozzle, performance variables were calculated for an ideal nozzle with

ideal expansion. We can model the nozzle based on solid propulsion methods since hybrid

propulsion methods are similar. Using mass flow rate and chamber pressures, we can

define and calculate performance variables such as the thrust coefficient (CF), Characteristic

Velocity (C*), Specific Impulse (Isp), Thrust Output (F), and Mach exit (Me). Design

restrictions were put into place to limit the chamber temperature (To) to 825 [K], which

allowed for calculating the ratio of specific heats (k) for the exhaust gas, which was

assumed to be CO2 because it is a known byproduct of all combustion reactions. A value for

specific heat was established to be 1.19125. The value of the Vandenkerckhove constant (Γ)

was calculated as 0.6468 using equation (7).

𝑘+1
2 2(𝑘−1)
Γ= 𝑘× 𝑘+1
(7)

The value for atmospheric pressure (Patm) was calculated for the future test site at Watkins

Space Port to be 83765.07 [Pa.]. The value for the max chamber pressure (Pch) was

6879021.278 [Pa.], calculated from the fuel grain as discussed above. Once all the input

variables were established, the values for the performance variables were calculated using

the following equations: equation (8) was for CF, equation (9) for C*, equation (10) for Isp,

equation (11) for F, and equation (12) for Me.

𝑘−1

2𝑘 𝑃𝑎𝑡𝑚 𝑘
𝐶𝐹 = Γ× 𝑘−1
×(1 − 𝑃𝑐ℎ
) (8)

* 1
𝐶 = Γ
× 𝑅×𝑇𝑜 (9)
*
𝐶𝐹×𝐶
𝐼𝑠𝑝 = 𝑔
(10)

*
𝐹 = 𝐶𝐹 × 𝐶 × 𝑚̇𝑡𝑜𝑡 (11)

(( )
𝑘−1
⎡ ⎤
𝑃𝑐ℎ
)
𝑘
2
𝑀𝑒 = ×⎢ − 1 ⎥ (12)
𝑘−1 ⎢ 𝑃𝑎𝑡𝑚 ⎥
⎣ ⎦

The performance variables were then used to calculate the throat area At using equation

(13) and the exit area Ae using equation (14).

*
𝐶 ×𝑚̇𝑡𝑜𝑡
𝐴𝑡 = 𝑃𝑐ℎ
(13)

𝑘+1

( ( )
)
𝑘−1 2 2(𝑘−1)
𝐴𝑡 1+ 2
×𝑀𝑒
𝐴𝑒 = 𝑀𝑒 𝑘+1 (14)
2

Using these two values, we are then able to calculate the throat diameter to be 0.62 [in.] and

the exit diameter of 2.2 [in.], using equations (15) and (16), respectively.

4𝐴𝑡
𝐷𝑡 = π
(15)

4𝐴𝑒
𝐷𝑒 = π
(16)
We can design an ideal nozzle defining all values while maintaining the calculated thrust

output of 273 [lbf.] and Mach exit of 3.28. Since our throat and exit diameters are more

significant than the previous designs, changes were made to the nozzle. The outside

diameter, length, and structure of the divergent/convergent sides of the nozzle were the

most significant changes in the design. Supersonic nozzles always consist of a convergent

section leading to a minimum area followed by a divergent section.

As seen in Figure 2.6.1 below, the divergent side of the nozzle maintained a conical design,

while the convergent side is now concave. The convergent side of the nozzle is a part of the

combustion chamber process. Computational Fluid Dynamics analysis was utilized, which

showed that we could improve our Mach exit to 3.48 compared to our calculated value of

3.28 while also dispersing ejected mass uniformly so that we do not experience an

overexpansion issue at lift-off. When an overexpanded flow passes through a nozzle, the

higher atmospheric pressure causes it to squeeze back inward and separate from the walls

of the nozzle. Thus, reducing the efficiency of generating any additional thrust. An

overexpansion scenario is standard at sea level because it corrects itself as the rocket gains

altitude. However, this is an issue we do not want to encounter since we are lifting off at a

higher altitude.

For the flow to become supersonic, certain conditions must be fulfilled. The ratio between

the inlet and exit pressures in all rocket propulsion systems must be sufficiently large to

induce supersonic flow. The throat of all ideal one-dimensional supersonic nozzles must

have a Mach number equal to one. Figure 2.1.5.1 shows that we can maintain around Mach

1 at the nozzle's throat and produce a supersonic flow at the exit while being dispersed as

evenly as possible.
Figure 2.1.5.1: CFD Analysis of Mach Number at exit for Nozzle

The nozzle will continue to be machined and manufactured from Carbon Graphite round

stock. Graphite is one of the most common materials used in rocket nozzles due to its

high-temperature resistance, low cost, machinability, high thermal conductivity, and

increasing toughness at elevated temperatures. This year, our diameters are larger for our

nozzle than the previous RocketLynx projects. Therefore, we will have to order a larger

diameter rod of Carbon Graphite to manufacture our nozzle.

In Figure 2.1.5.2, a stress analysis was conducted to predict what stresses the nozzle will

experience and make sure it is within the limits of the Carbon Graphite. The figure shows

that the highest von Mises stress experienced is 2.4251007[N/m2], around the convergent

side of the nozzle where it is part of the combustion chamber. The total yield strength for

Carbon Graphite is 1.2061008[N/m2], which shows we are within limits and deformation is

not expected.
Figure 2.1.5.2: Stress Analysis of Nozzle

After conducting further analysis on the nozzle, continuing on from the Fall 2022 iteration,

we noticed that over expansion was still an issue with the nozzle. We’ve decided to use the

rear flange of the combustion chamber to assist the nozzle with over expansion at the exit.

The final dimensions of the nozzle are displayed in table 2.1.5.1:

Table (2.1.5.1): Nozzle Dimensions


Dimensions of Nozzle
Dimensions
Outside Diameter = 3.46 [in]
Length = 4.12 [in]
Convergent Diameter = 2.98 [in]
Convergent Angle = 55°
Divergent Diameter = 2.20 [in]
Divergent Angle = 241°

Throat Diameter = 0.62 [in]


O-Ring Gland Width = 019 [in]
O-Ring Gland Depth = 0.12 [in]

Distance Between O-Ring = 0.11 [in]


Glands = 2
The dimensions displayed in Table (2.1.5.1) were used to manufacture the nozzle. The Trax

Lathe was used to conduct the Facing, Drilling, Boring, and Grooving operations to

manufacture the nozzle. Safety precautions were also set in place while manufacturing the

nozzle. We wore protective clothing, N95 respirator masks, and held a vacuum hose near

the nozzle while it was being manufactured, to ensure as much of the carbon graphite dust

particles were removed from the immediate area. This was done because this material is

electrically conductive and could have shorted out the machines if too much build up

occurs. Also the dust particles produced can irritate our respiratory systems, hence the

reason to wear respiratory protection masks.


2.2. Oxidizer Team

2.2.1. Pressure Vessel Current State (Sean and Garrett)

Since the midterm report in Spring 2023, the bolt holes were drilled in the pressure vessel.

This was the very last machining operation performed on the pressure vessel. It was fitted

with all hardware, consisting of Buna-N o-rings (dash no. 258), as well as ¼ [in] fine thread

grade 8 steel bolts, nuts, and split lock washers. The pressure vessel was outfitted with

plumbing necessary for the hydrostatic test, and then the test was performed successfully

up to approximately 1050 [psi]. From this point forward, the pressure vessel was

considered to be completely finished and performing as expected from the design. The

pressure vessel was outfitted with correct plumbing needed for test day, and performed as

expected for the live test fire.

The only issue present during test day was that the team was unable to fill the pressure

vessel to the desired capacity. It was only filled with about 7 [lb] of nitrous, and only

pressurized to about 600 [psi]. The goal was to have 14 [lb] of nitrous, an internal pressure

of approximately 800 [psi]. This was not a limitation caused by the design of the pressure

vessel.
2.2.2. Injector Current State (Ryan and Luis)

The injector plays a critical role in the efficiency of the burning of the solid fuel grain

by atomizing the N2O from the run system into the combustion chamber. Atomization is the

process of converting a liquid into exceedingly small droplets and is the primary

consideration when designing an injector. The size of the droplets directly affects how

efficiently and rapidly combustion of the fuel grain will occur. The small droplet sizing

causes the oxidizer to behave more as a gas than it does as a liquid. This process referred to

as “Atomization” occurs when the appropriate Ohnsorge (Oh) to Reynolds number (Re)

ratio is implemented. The variables of interest are fluid density (ρ), flow speed (v), orifice

diameter (Dorifice), dynamic viscosity (μ), and Weber number (We). The Re and Oh were

calculated using equations (29) and (30).

ρ𝑣𝐷𝑜𝑟𝑓𝑖𝑐𝑒
𝑅𝑒 = µ
(29)

𝑊𝑒
𝑜ℎ = 𝑅𝑒
(30)

The calculated Reynolds number and Ohnsorge number were 865972 and .001002,

−9
respectively. Using these values, the Oh/Re ratio was calculated to be 1. 1574710×10 .

Figure 2.2.2.1 was used to determine atomization


Figure 2.2.2.1: Ohnesorge-Reynolds Plot

Past iterations of the design used a single port or shower head design. This year's team

decided on a modified showerhead design with an increased number of holes as well as

impinging exits to improve performance. Based on findings of similar propulsion rocket

designs, the diameter of the injector plate as well as orifice diameter were determined to be

0.0127 [m] and 1.5 [mm], respectively. Equation (31) was used to compute the number of

orifices necessary in the injector for optimal atomization. The variables of interest are

oxidizer flow rate (ṁ), area of orifice (Aorifices), change in pressure (P), density of oxidizer

(ρox), and energy loss constant (K).

𝑚̇ 5
𝑁𝑜𝑟𝑓𝑖𝑐𝑒𝑠 = 𝐴𝑜𝑟𝑓𝑖𝑐𝑒𝑠 2ρ(𝑁2𝑂)∆𝑃
(31)
The number of orifices were calculated to be 12.7 which we rounded up to 13 to provide

symmetry. The droplet sizes were determined using equation (32) to be 0.08 [mm]. The

variables of interest are injector diameter (Dinj), surface tension(σ), density of oxidizer (ρ),

viscosity (μ), flow velocity (v), and gravity (g).

( )
1
𝐷𝑖𝑛𝑗 ⎡ ⎤
𝑆𝑀𝐷 = 47 𝑢 ( )
σ
ρ×𝑔
4
⎢1 + 331

𝑣
(ρ×σ×𝐷𝑖𝑛𝑗)
1
2


(32)
⎣ ⎦

In the design process of the injector, when deciding the positioning of the orifices, the fuel

grain shape was taken into consideration. Using a straight port design, 5 centrally located

orifices were placed on the injector plate to travel through the straight port. The following 8

outer diameter orifices will be the impinging to create further atomization of the oxidizer

amongst the fuel grain. Figure 3.5.2 was the design for the injector reflecting the findings

and concepts discussed.

Figure 2.2.2.2: Current Injector Design


It is important when designing the injector that the pressures and velocity at the inlet of the

injector are greater than the expected pressures in the combustion chamber; this will

ensure a steady flow of oxidizer as well as prevent backflow into the run system.

Computational Fluid Dynamics (CFD) Flow simulations were run on the injector to verify

adequate velocity and pressure are present when the oxidizer reaches the combustion

chamber.

Figure 2.2.2.3: CFD Velocity Simulation of Injector Design


Figure 2.2.2.4: CFD Pressure Simulation of Injector Design

Figure 2.2.2.3 and Figure 2.2.2.4 show that the injector design will generate the desired

pressure and velocity to compensate for the forces expected within the combustion

chamber.
2.3. Avionics Team

2.3.1. Master On/Off Current State (Alex)

After initial testing of the master on/off control system, a new housing unit needed to be

designed to incorporate the circuit used to control stopping or starting flow into the

combustion chamber. The master on/off control system will be following the team

throughout all of testing and thus must be portable and reliable.

The design of the housing unit consists of a lower half and upper half that will come

together to act as an enclosure for the circuit as shown in Figure 2.3.1.1. The circuit itself

will be soldered onto a permanent breadboard which will live in the bottom half of the

housing unit. The top half will have inserts for the buttons and LEDs to sit while also having

openings for the wiring needed to connect the LEDs, buttons, and servo. The two halves will

be put together using a M4 threaded insert and four M4 screws. New buttons have been

implemented into this design to increase reliability and functionality.

Figure 2.3.1.1: Master On/Off Housing


2.3.2. Ignition Circuit Redesign (Alex)

Initial testing of the igniter circuit was previously completed and proved to be successful.

However, as the team began to move the circuit into a more reliable orientation, the circuit

began to fail in the early Spring 2023 semester. After doing some trouble shooting, the team

was able to narrow the problem to low battery source, low amperage, and wiring

compromise.

The first step to solving this problem was checking the battery source. Using an altimeter,

the team was able to determine that the battery was producing 8.5 [V] of voltage to the

circuit. Following this, the team decided to check where the voltage was traveling through.

While doing this, it was found that about 6.3 [V] were going into the first igniter wire and

only 2.2 [V] were going through the second igniter wire. While the voltage was not

spreading evenly, it should not have compromised the circuit in any way. The team decided

to focus on the amperage going through the circuit next. The Firewire Initiator being used

requires 0.6 [amps] for ignition and thus gave a team an idea of where an issue may have

been lying. By using the altimeter, the team determined that the resistors used for the LED’s

and the Firewire acting as a resistor itself, the amperage on the second ignition wire was

being cut to just 0.1 [amps]. In order to solve this issue, the team looked at adding a second

battery connected in series to increase the amps through the circuit. By doing this, the team

had increased the voltage to roughly 16 [V]. As this was tested, it was found that the

ignition wire was still not igniting. The team then decided to change the orientation of the

circuit from parallel to series. This would allow the same amperage to pass through both

the igniter wires. While initial testing looked promising, reliable results were still not being

achieved. As a last ditch effort, it was decided to use two brand new batteries in series to
determine if there was an issue with the power source. This showed the team that the two

batteries being used were old and running out of voltage quickly. Using two new batteries

provided reliable ignition for multiple tests. For this reason, the team has decided to use

two brand new batteries for the day of testing.

With the issues being solved for the ignition circuit, the next steps are to begin soldering

the circuit onto a permanent breadboard along with enclosing it into the housing unit.

Figure 5.1.1 shows the housing unit design which was 3D printed due to the ease of

manufacturing. The housing unit will have the key switch and button placed into it where it

will be soldered directly onto the permanent breadboard. Along with this, a red LED and a

green LED will be connected to show the team when the circuit is ready for ignition as well

as when the circuit has fired.

Figure 2.3.2.1: Ignition Circuit Housing Unit


2.4. Test Stand Redesign (Alex)

The test stand designed by previous iterations of the team was made to hold the rocket

motor in place during static testing while also collecting data on the thrust output of the

motor during each test. This year’s iteration of the team will redesign a new test stand that

can be manufactured from the existing test stand that is available to us. The base stand is

made of steel square tubing welded in a configuration that gives support to the combustion

chamber. The new design of the test stand takes this configuration and adds to it. First, the

team has decided to extend the test stand by 18 [in]. to allow clearance for the new

combustion chamber as well as the run system. With the test stand being extended by 18

[in], the team also has the option to mount the new pressure vessel. At the extension, the

team will weld on new steel tubing in a vertical orientation for mounting the pressure

vessel. This is being done to ensure the pressure vessel has the best chance of releasing

nitrous oxide at the mass flow rate needed for the combustion chamber. The vertical tubing

will also have supports welded on at a 45 degree angle to ensure there is no failure in the

test stand. Finally, the team will weld a small extension in the middle of the frame, where

the combustion chamber begins, to allow for mounting of the load cell. This will allow for

about 90% of the load cell to hang off with about 0.6 in of ground clearance to the floor.

Finite element analysis (FEA) was performed on this newly designed test stand to ensure

the test stand will have no failure during the static hot-fire tests. Within the FEA, the

bottom of the stand was fixed in place to mimic its interaction with the ground. This also

prevents translational and rotational movement of the test stand itself. After this, bolt

connections with a load on them were incorporated to imitate the weight of the pressure

vessel loaded with nitrous oxide. Finally, a distributed load was applied on the extension to
imitate the thrust output of the motor going onto the load cell. The distributed load was set

to 275 [lbf] which is the maximum thrust output the team is expecting to see out of the

motor. All the component material in the simulation was selected as steel: AISI 1020.

Figure 2.4.1: Test Stand FEA Von-Mises Stress Plot

Based on the FEA results shown in Figure 2.4.1, the maximum Von-Mises stress is about

16.3 [MPa] which is far less than the maximum yield strength of the material at 351.6

[MPa]. This max stress can be seen being applied directly onto the extension where the load

cell will be held but will not cause any failure at that point. Based on this analysis, the test

stand will be able to withstand all the forces being applied to it from both the motor and the

pressure vessel. With this analysis giving promising results, the team will be able to move

forward with manufacturing and assembly.


3. Manufacturing Methods (Alex)

The essential goal for this semester was to allow each team member to get hands-on

manufacturing experience. The following sections discuss the planning, process, challenges,

and results for the manufacturing of each engine component. For manufacturing to be

completed efficiently, the team was split into teams where each team worked on specific

components. The teams were the propulsion team and the oxidizer team. Along with those

two teams working on their specific components, the team has decided to work on the test

stand as a collective unit. With each component being managed by different members of the

team, the team was able to share knowledge across the team. The knowledge shared

involved computer-aided manufacturing (CAM), machining, additive manufacturing, and

best practices for bettering processes. This allowed team members to realize the

importance of their design and analysis throughout the project.

3.1. Propulsion Team

3.1.1. Fuel Grain Manufacturing (Marco and Willis)

The fuel grain was manufactured using the Original Prusa i3 MK3 printer (Figure 3.1.1.1).

This device has a nozzle of 0.8 [mm.] These features helped to print the fuel grain faster and

increased the strength of the material. The printing resolution of the fuel grain was 0.6

[mm.], and the infill was approximately 70%. As mentioned, the fuel grain was printed in

three parts since its length increased, and the printer could only operate within a specific

range. A simple interlocking key system was designed to assemble the three parts.
Figure 3.1.1.1: Original Prusa MK3 Printer

Lastly, ABS cement was applied to all mating faces of the fuel grain, clamped, and allowed to

cure for 24 hours before they were inserted into the combustion chamber (Figure 3.1.1.2).

Figure 3.1.1.2: Assembly of fuel grain using ABS cement

3.1.2. Combustion Chamber Manufacturing (Erin and Michael)


Due to manufacturing limitations, we decided to manufacture the combustion chamber in

two parts rather than one. The overall chamber length of 25.5 [in] was too long to

manufacture on the HAAS TL1 lathe in the Hub. The option of sending the part out to be

manufactured was ruled out due to budget constraints. The two new pieces were designed

to be exactly half of the original design at 12.75 [in] each to limit the number of calculations

we would need to redo, given the design change. We included a counterbore design to

ensure these two pieces fit together with limited internal combustion and flow disturbance.

The counter bore was manufactured with the utmost care to ensure limited dimension

deviation so the two pieces would fit snugly together. The parts were also held together

using flanges and bolts similar to the outer flanges previously developed for the overall

combustion design. The combustion chamber outer shell was separated from the heat and

flow with our heat-resistant liner and the fuel grain, so the effect of any gap or lip was

minimal.

We finished most of the manufacturing on the combustion chamber with all our measured

dimensions being within 0.05 [in] of our original design. This discrepancy is minimal and

did not affect performance. While most of the manufacturing time was spent on the lathe,

we also used a bandsaw. The band saw cut the 6 [ft] aluminum stock into pieces with the

rough dimension of the two combustion chamber pieces plus 1 ¾ [in] to ensure enough

material for facing and the chucks on the lather to grab. We then used the lathe to cut the

inner and outer diameters.

The outer diameter was largely manufactured by a right-hand tool. However, a groove tool

was initially used to create a gap big enough for the right-hand tool to fit in. A finishing pass
was also taken over the outer diameter to ensure a smooth finish. The inner diameter was

manufactured using two ID boring bars. A set of 0.700 [in] and 0.875 [in] drill bits were

used for starting holes big enough for the boring bars to fit in the combustion chamber to

carve the interior diameter. Again, a finishing pass was taken to ensure a smooth finish. The

boring bars were only long enough to fit about ¾ of the way down the pieces, so we

manufactured half of the inner diameter, then flipped the piece and manufactured the other

half with a ¼ [in] overlap.

We used a feed rate of 0.012 [in/rpm] for the bulk of the material removal but lowered the

feed rate to 0.006 [in/rpm] for the finishing passes to have a smooth finish. Our speed on

the mill varied from 100 [rpm] to 300 [rpm] depending on the tool and the distance cutting

from the chucks. Finally, we finished the manufacturing process on the combustion

chamber by using the mill to drill the bolt holes. The diameter of the bolt holes was

calculated to be 9/16 [in]. Once these calculations were complete and verified, the mill was

used to bore out eight identical, evenly spaced holes. An overview of the CAM process can

be seen below in Figure 3.1.2.1.


Figure 3.1.2.1: CAM of the Combustion Chamber
3.1.3. Nozzle Manufacturing (Joel)

The graphite rod stock was cut with a Bandsaw to an approximate length of 4.5 [in.] to fit

inside the lathe spindle. Once the stock was in the lathe, a turning operation was conducted

to cut the stock down to the required outside diameter. Figure 3.1.3.1 displays the CAM

operation currently being used for this operation.

Figure 3.1.3.1: Cam Simulation of Outside Profile of Nozzle

Once the turning operation was completed, the facing operation was conducted. This

ensures we have even surfaces on both sides of the stock to conduct the rest of the

operations. Figure 3.1.3.2 shows one completed side of the facing operation..
Figure 3.1.3.2: Facing Operations of Nozzle

Next, the drilling operation was conducted. Instead of doing two drilling operations for the

throat diameter through the entire nozzle, like originally planned, we used a 15/32 drill bit

to deep drill through the entire length. Then we allowed the boring operations to cut the

diameters to the correct sizes.

To manufacture the convergent and divergent sides of the nozzle, a boring operation was

done with the lathe. Standard-size tooling was used for these procedures. Figures 3.1.3.3
and 3.1.3.4 respectively, display the results for the boring procedures of the convergent and

divergent sections of the nozzle.

Figure 3.1.3.3: Boring of Nozzle Convergent Section

Figure 3.1.3.4: Boring of Nozzle Divergent Section


Manufacturing the glands for the O-rings was the last operation. The O-ring glands were

designed to use -236 PTFE O-rings. Figure 3.1.3.5 displays the final operation results.

Figure 3.1.3.5: O-Ring Glands Operation

Manufacturing the nozzle required safety precautions to be in place. Carbon graphite can

affect the respiratory system if inhaled. Irritations can occur if it comes into contact with

the eyes or skin. Carbon graphite will produce a good amount of electrically conductive

dust particles, and coolant cannot be used with the lathe while manufacturing this material.

Respiratory protection masks, safety glasses, protective clothing, and adequate ventilation

systems were used while manufacturing the carbon graphite nozzle. A vacuum was used
while manufacturing the nozzle to ensure the machines in the immediate vicinity were not

shorted out.

3.1.4. Nozzle Flange (Erin)

The nozzle flange was manufactured out of a square 1 ¼ [in] thick Aluminum 6061-T6 slab.

The manufacturing for this component was mostly done on the mill, the center hole was

bored on the mill with a 1[in] boring tool, center holes were done with a #3 center drill,

bolt holes were drilled using a 9/16 [in] drill and the piece was faced on the lathe to 1.2 [in]

thickness.
3.2. Oxidizer Team

3.2.1. Pressure Vessel Manufacturing (Sean and Garrett)

Manufacturing of all pressure vessel components has been completed. To machine the 17 [in] x 7.5

[in] body of the pressure vessel, a piece of aluminum 6061 T-6 stock was cut to 18.25 [in] x 7.5 [in].

The additional 1.25 [in] of material was added to allow for facing operations as well as providing

sufficient material for the lathe chuck to secure the part. Exterior turning operations began by

using a dial indicator to center the part on the lathe and drilling a No.8 center drill into the

face of the part. This center drill hole allowed for the use of a live tailstock to secure the end

of the part. A right-hand tool was then used to conduct a facing operation on the exterior of

the part. This was done to ensure a square face and cylindrical body. Next, a grooving tool

was used to machine a 0.75 [in] flange into the exterior of the aluminum stock. This was

accomplished by cutting a 0.5 [in] deep x 0.12 [in] wide groove and then moving laterally

one tool width and repeating the process. This continued until a total groove width of 16

[in] was achieved. An additional iteration was completed, cutting to a depth of 0.25 [in], to

reach the final flange height of 0.75 [in]. The grooving process can be seen in Figure 3.2.1.1.
Figure 3.2.1.1: OD Grooving Operation

The gland for the O-ring was then machined into the face of the part. This was done using a

custom fabricated 0.1480 [in] wide grooving tool. The Haas lathe does not have a built-in

face grooving operation, so a CAM program was created using Autodesk Fusion 360. This

allowed us to ensure a groove depth of 0.11 [in] and width of 0.16 [in] would be achieved to

accommodate the required O-ring. Boring the ID of the pressure vessel began by manual

drilling operations on the lathe to create a clearance hole for the boring bars. First, a 0.70

[in] drill was used to create a pilot hole half the length of the part,

or approximately 9 [in]. A 0.875 [in] drill was then used to create adequate clearance for

the initial 0.726 [in] wide boring bar. Drilling operations can be seen in Figure 3.2.1.2.

Figure 3.2.1.2: ID Drilling Operation

Another CAM program was created and utilized to machine an 8.518 [in] deep x 1.3 [in]

diameter hole to allow clearance for the larger 1.11 [in] diameter boring bar. The larger

boring bar was then used to bore the stock to the final ID of 5.5 [in] using a similar CAM

process as the small boring bar. The ID boring operations can be seen in Figure 3.2.1.3.
Figure 3.2.1.3: ID Boring Operation

Due to the symmetry of the part, the part was able to be flipped, and the same process was

completed on the opposite side. The machined pressure vessel body can be seen in Figure

3.2.1.4.

Figure 3.2.1.4: Pressure Vessel Body


Two identical hemispherical end caps were each machined out of a 4.5 [in] piece of

aluminum 6061 T-6 stock. First, a piece of aluminum stock was centered and located on the

Trak Lathe. The interior was then bored to a 2.75 [in] radius using an ID Boring bar. This

can be seen in Figure 3.2.1.5 below.

Figure 3.2.1.5: End Cap ID Boring Operation

The part was then flipped on the lathe, and the exterior was machined using a right-hand

tool which can be seen in Figure 3.2.1.6.

Figure 3.2.1.6: End Cap OD Machining


Next, threads for the inlet/exit were machined into the endcap. This was done by first

drilling a 7/16 [in] hole into the inlet/exit, and then using a ¼ [in] - 18 NPT tap to cut the

threads. A ¼ [in] - 18 NPT pipe threaded plug gauge was used to ensure threads were cut

to the proper depth. Use of the plug gauge can be seen in Figure 3.2.1.7.

Figure 3.2.1.7: Measuring End Cap Thread Depth

Eight evenly spaced bolt holes were then drilled into each flange of the pressure vessel

body, as well as the flange on each end cap. Drilling operations for all components were

conducted in the same manner on the Vectrax mill. First, each part was located on the mill

using an edge finder as seen in Figure 3.2.1.8.


Figure 3.2.1.8: Locating Part Using Edge Finder

Once the part was located, the origin was set at the center of the part. X and Y coordinates

for each hole were then taken from the Solidworks CAD model. Drilling operations began by

manually traversing to the X and Y coordinate of each hole and using a ¼ [in] center drill to

create a pilot hole. To accommodate ¼ bolts, an H drill bit (0.266 [in]) was used to allow for

adequate clearance of the bolt. The same process used for the center drill was then

repeated using the large H drill to bore each hole to the final diameter. Drilling of the bolt

holes can be seen in Figure 3.2.1.9 below.

Figure 3.2.1.9: Drilling Flange Bolt Holes.

The final step was the assembly of the pressure vessel. After inserting Buna-N O-rings (dash

no. 258, 90A Durometer) into each gland on the end of the pressure vessel body, ¼ [in]

diameter by 1¾ [in] long 28 UNF grade 8 bolts and nuts along with split lock washers were

used to secure the end caps to the main body. All bolts were torqued to 15 [lb/ft] of torque

in a star pattern to ensure even mating of the surfaces. The completed pressure vessel

assembly can be seen in Figure 3.2.1.10.


Figure 3.2.1.10: Complete Pressure Vessel Assembly

3.2.2. Injector Manufacturing (Ryan and Luis)

Manufacturing of the injector has been completed. The current manufacturing process for

the injector was done using a lathe, with tool changes being conducted depending on the

operation. Identifying the right stock and creating the necessary CAM operations were the

first items to complete. The injector was initially designed and analyzed to be made out of

aluminum 6061-T6, but the stock material was later changed to stainless steel. Due to

complications related to drilling the team decided to switch the material back to aluminum

6061-T6. Two additional injectors were manufactured until the final injector was

completed. The additional injectors were results of issues with tolerancing and structural

damages. The injectors largest OD was designed to be d= 1.28 inch and overall length L= .69
inch. It was determined that the best approach would be to machine the injector from its

back side. Utilizing the Trak TRL 1630SX Lathe, the injector stock was first faced, then

followed by a outside profile roughing and finishing using a #1 Right hand tool with a

corner radius of 0.03125 inch. To separate the stock from the injectors outside profile, a

grooving operation was done using a #8 OD grooving tool with a corner radius of 0 inch. In

order to accomplish successful atomization 13 holes were drilled through both faces of the

injector. Five non-impinging inner diameter holes were completed utilizing the Haas mill.

Eight impinging outer diameter holes were completed utilizing a dividing head that was

attached to the haas mill. The fuel injector was successfully manufactured using the stated

operations and the final result is shown below.

Figure 3.2.2.1: Back Face of Injector


Figure 3.2.2.2: Front Face of Injector

3.2.3. Injector End Cap Manufacturing (Ryan)

The End Cap will be made from a 7.5 [in] diameter 6061 round aluminum stock that will

require a turning operation to get down to the necessary diameter of 7 [in]. Due to the need

for both turning and milling operations, determinations were made in the initial stock

length of 4 [in] to optimize the set-up process and expedite machining time. It was

determined that the best approach for the process was to machine the top (oxidizer side) of

the end cap first.  Utilizing the Trak TRL 1630SX Lathe, the end cap was first faced, then an

outside profile roughing and finishing using a #1 Right hand tool with a corner radius of

0.03125 [in]. Following the completion of the outside profile, the utilization of a 0.5 [in]

center drill was used to start the initial central hole of the end cap. This operation was

followed by a 0.5 [in] drilling operation all the way through the stock. This was followed by

a 0.75 [in] drilling operation throughout the entirety of the end cap. This created a starting

point to work from for our central hole, which will have a final minimum diameter of 0.84
[in]. After the initial drilling operations were complete, using Autodesk Fusion, a boring

operation was set up to bore the top half of the end cap. This boring operation will be

where the injector is seated and is shown in Figure 3.2.3.1 Using a #4 boring bar, an inner

profile roughing and finishing was completed to create the necessary stepped inner

diameters of 0.86 [in], 1 [in], and 1.28 [in] respectively, required for the injector to sit flush

inside the end cap. 

Figure 3.2.3.1: Boring Operation of Injector side (Top)

Upon completion of the top region of the end cap, the stock was flipped, and the machine

was recalibrated for accurate X and Z positioning as well as height, to ensure accuracy of

the machining. Like the top operations, the bottom of the end cap will be faced, followed by

roughing, and finishing operation using the same #1 right hand tool. The turning

operations of the bottom half of the centrally located hole will utilize an inside turning

operation using a #4 boring bar to create a conical shaped nozzle angled at 54.98° outward

as shown in Figure 3.2.3.2. The exterior wall of the nozzle shape is where the endcap will
meet the combustion chamber, therefore glands were made for O-rings on the exterior

diameter using the #8 Grooving tool that has a corner radius of 0.003 [in]. The part will

then be transferred to the mill where the first operation is to face the excess material on the

part that was being used to clamp the part with a 2 [in] flat end mill. The eight bolt holes

are then drilled with a 5/16 [in] drill, and the stepped IDs on the side of the part are then

milled with a ⅜ [in] flat end mill. The drilled bolt holes will be renamed with a 5/16 [in]

reamer and threaded manually to an M5 specification as those are bolts that were selected

for this application.

Figure 3.2.3.2: Boring Operation of Combustion Chamber Side (Bottom)

The End cap was successfully manufactured using the stated operations and the final result

is shown below.
Figure 3.2.3.2: Completed Injector End Cap

3.2.4. Oxidizer Cap Manufacturing (Luis)

Manufacturing for the oxidizer cap has been completed. The manufacturing process for the

oxidizer cap was done using a lathe and mill machine, with tool changes being conducted

depending on the operation. Identifying the right stock and creating the necessary CAM

operations were the first items to complete. The oxidizer cap was designed and analyzed to

be made out of aluminum 6061-T6. The oxidizer caps largest OD was designed to be d= 7.5

inch and its smallest ID to be d=.84 inch. The overall length will be L= 3.25 inch. Utilizing

the Trak TRL 1630SX Lathe, the oxidizer cap was first faced, then followed by a outside

profile roughing and finishing using a #1 left hand tool with a corner radius of 0.2 mm.
Figure 3.2.4.1: Turning Operation (Machine OD profile)

Following the completion of the outside profile a 7/16 inch center drill was used to center

drill the face with the tapered holes. Following the center drill operation a 45/64 inch drill

was used to drill the tapered hole side. The stock was then flipped and utilizing the Trak

TRL 1630SX Lathe the top face was faced using a #1 left hand tool with a corner radius of

0.2 mm. Following the completion of the facing operation a 7/16 inch center drill was used

to center drill the top face. Following the center drill operation, a ⅞ inch drill was used to

drill the top face. A tap and die will then be used to cut tapered threads that measure ½

inch x 14 NPT.

Figure 3.2.4.2: Profile Roughing Oxidizer Cap (Tapered Hole Side)

The Oxidizer Cap was successfully manufactured using the stated operations and the final

result is shown below.


Figure 3.2.4.3: Finished Oxidizer Cap

4. Test Stand Manufacturing (Joel,Erin)

To update the RocketLynx IV test stand in order for it to work for RocktLynx V’s design a

few components needed to be added. In order to add these components Gas Metal Arc

welding was used to secure uprights, side supports, D-rings, combustion chamber cages,

and additional length. This was done according to the thickness of the steel square tubing

used. The two thicknesses that were used were 12 and 14 gauge tubing. The settings of the

brazing/MIG welder were adjusted to the necessary feed rate and voltage based off of the

largest thicknesses being connected to create proper penetration of the metal without

compromising the integrity of the tubing or the weld. Proper PPE was worn during all

welding. This included wearing 100% cotton with no exposed skin on the arms or legs,

welding helmets, leather gloves, and steel toed boots.

5. Avionics System Manufacturing

5.1. Ignitor Circuit (Alex)


Manufacturing of the ignitor circuit required soldering, 3D printing, heat shrinking, and

finally ended in assembly. The first step in manufacturing the ignition circuit was soldering

all resistors, wires, diodes, and battery power/grounds onto a permanent breadboard. This

was done with the soldering iron set to 600 degrees Fahrenheit. To ensure the circuit was

being soldered correctly, a schematic along with an example circuit was available for

reference during the soldering process. Along with this, the team ensured to inspect each

solder to ensure that proper fill was being achieved within each soldering point. To solder

the LEDs, key switch, and button, extension wires were soldered directly to their

connection points so that they can sit snugly in the housing unit. While the soldering

process was taking place, 3D printing of the housing unit was being done. The housing itself

was printed on a Prusa i3 MK3 printer utilizing PLA plastic. The infill for the housing unit

was decided to be 20% due to the high strength and light weightiness this infill produces.

Once the housing unit was printed and all soldering was done, the next step was to insert

the M4 threaded inserts. The M4 threaded inserts were installed utilizing the soldering iron

and the ability to melt into the PLA plastic which makes up the housing unit. After installing

the M4 threaded inserts, the team placed the key switch, button, and LED’s into the housing

unit. Once these critical components were installed, the batteries and permanent

breadboard were all placed into their designated spots. Finally, to close the entire assembly,

M4 screws were utilized to close the entire assembly. During this process, it was realized

that the clearance required for the M4 screws was not achieved. The team decided to use a

power drill with an M5 sized drill to cut out enough clearance for the head of the M4 screw.

This allowed the M4 screw to fit directly into the housing unit and allowed for the circuit to

be closed up. The final result of the housing unit can be seen in Figure 5.1.1.
Figure 5.1.1: Igniter Housing Unit

5.2. Master On/off (Alex)

The master On/Off circuit followed very similar manufacturing processes as the ignition

circuit. In order to manufacture the master on/off circuit, soldering, 3D printing, and

assembly was all required for the final product. The first step in manufacturing the master

on/off circuit was soldering all resistors, button connections, LEDs, and power/ground

connections onto a permanent breadboard. Once again, a schematic as well as an example

circuit was provided during the soldering process to ensure all solder points were correct

for the circuit. The team also ensured to check all solder points for appropriate solder

filling. This was to ensure that connections were secure but also making contact with the

leads of the circuit. While the soldering process was underway, 3D printing was once again

utilized to print the housing unit of the circuit. The housing unit was printed on the Prusa i3

MK3 printer with an infill ratio of 20%. For the master on/off circuit, the M4 threaded

inserts were inserted first before inserting the entire circuit. Once all main connections
were soldered onto the permanent breadboard, extension wires were added to the LEDs

and buttons and then soldered into their location on the circuit. Once all points were

soldered together with the LEDs and buttons placed into their housing unit, the housing

unit was closed up utilizing M4 screws. With the housing unit for the master on/off circuit

complete, the final step was to solder on 200 feet of wire onto the LEDs and buttons

extension wires. With the arduino used to control the master on/off being on the avionics

module next to the rocket system, 200 feet of wire was used to connect the housing unit to

the arduino. The final assembly of the master on/off can be seen in Figure 5.2.2.

Figure 5.2.2: Master On/Off Housing Unit


6. Test Preparations (Alex)

The team is currently having discussions with Bandimere Speedway in order to secure a

test site as well as a test date that will work with all parties involved. Bandimere Speedway

will be able to provide a large open lot to protect the team and the environment from

possible misfires and eruption of components. Along with that, the team is planning safety

procedures for all parties involved in order to keep all members safe but also to ensure the

team is prepared for emergencies.

Prior to testing, the team must complete an inspection on all manufactured components to

ensure each part was manufactured within acceptable tolerances. The combustion chamber

is of utmost importance as it was designed in two pieces where the two parts will come

together to interlock with one another. Ensuring that tolerances were met will ensure that

there is no leaking of the oxidizer or exhaust gas while the combustion process is occurring.

Possible leaking may cause a decrease in the motor’s performance resulting in lower thrust

output than expected. The combustion chamber as a whole will be secured in the HUB to

ensure that all bolts are torqued down to the required torque. This will ensure that tools

are readily available to the team for the combustion chamber and will allow the team to

transfer the chamber in its full assembly. Assembling the combustion chamber prior to

testing will allow for a quicker assembly on site during test day.

The next step will be to test for the oxidizer system prior to a full-scale test fire of the

motor. The team will first need to assemble the run system, consisting of pipes, valves, and

control systems which all lead from the oxidizer tank to the injector and into the

combustion chamber. This will ensure that all parts of the system fit correctly and will

allow the team to run a cold flow test. Cold flow testing will be done by utilizing the master
on/off circuit in order to test the system’s ability to activate the flow of the oxidizer as well

as properly atomize the oxidizer as it moves through the exit of the injector. The cold flow

test will also allow the team to check for any leaks in a safe manner. Along with testing the

atomization, the team will be able to determine the reliability and functionality of the

control system present.

The next test that will need to be done will include the pressure vessel and feed system. The

pressure vessel will need to go through a hydrostatic pressure test to determine the safety

of the newly designed tank. Putting the pressure vessel through a hydrostatic pressure test

will allow the team to look for leaks in the tank as well as ensure the team can fill it in a safe

manner. This will be an important test as the team plans to fill the pressure vessel with

nitrous oxide at the test site for each test fire that is done. After hydrostatic testing of the

pressure vessel is done, the team will begin testing the feed system to determine the

feasibility of filling the system using gravity feed. The feed system will be fully assembled

and connected to the pressure vessel. The team will be able to utilize the 10 [lb] nitrous

oxide bottle available to us to begin testing the filling system. This test will also allow us to

check for any leaks and ensure all components are working as intended. Finally, this test

will let the team know if the bleeder valve works as intended to ensure the feed system

does not get over pressurized.

The avionics team will be working towards completing all final wiring for all systems to

ensure the team is able to stand at the required distance of 200 feet. The wiring of the full

system will be done with considerations of a quick setup and tear down of all control

systems being used on test day. This will require the team to solder wire connections and

permanent connections to the appropriate circuit boards being utilized. Once all control
systems are wired, the team can continue testing of all control systems with cold flow tests,

loading tests, and safety checks.

6.1. Feed System (Ryan, Alex)

The purpose of the feed system is to transfer N2O from the mother (supply) tank to the

pressure vessel. The feed system has three major components to consider in the design

process. The first is the properties of Nitrous Oxide. With operation pressures anticipated

to be in the range of 750-800 [psia.], it is important to consider the material and pressure

ratings of components in the feed system. N2O at 70 [℉] has a pressure of 760 [psia.] and

approximately 1000 [psia.] at temperatures over 100 [℉]. To be able to sustain these

pressures, all components of the feed system will have a maximum pressure rating of 3000

[psi.] or greater to ensure an adequate safety factor and reliability. Therefore, high-pressure

stainless-steel piping and associated tee fittings were used in the connection of each

component. The second consideration is the effectiveness of the design. The feed system

must quickly and efficiently fill the pressure vessel to facilitate iterative test fires. To

achieve this, rather than using a Hy-Lok Q type quick disconnect that was previously

discussed. It was determined that simply utilizing the bleeder valve in order to

depressurize the system and using a standard N2O supply line would not only work just as

well, but would also be cheaper, and would allow more mobility of the mother tank while

filling due to the N2O line being a braided hose line which is substantially less rigid than  the

steel Hy-Lok quick disconnect. The third consideration, and the most important, is safety.

To ensure the safety of the operators, installation of ball valves on both the supply tank and

pressure vessel was implemented to isolate the N2O to its respective areas. A 500-1500
[psia.] rated N2O safety relief valve was installed onto the pressure vessel. The Safety relief

valve was set to 1000 [psia.] and calibrated using the appropriate gauges. Another safety

implementation that was made is a bleeder valve located in between the ball valves and the

disconnection point. It is important before disconnecting the feed system to relieve the

pressure in the piping. This will mitigate the risk of injury to the operator when

disconnecting the system.

Figure 6.1.1: Block Diagram for Feed Subsystem

Figure 6.1.1 displays how all three considerations were implemented in the design of the

feed system shown below was the final Result of the feed system.
Figure 6.1.2: Setup of Feed System

Shown in Figure (6.1.2) are: Prototype Pressure Vessel (Red), SPRV Pressure Relief Valve

(Blue), Shut Off Valve(Green), Bleeder Valve (Orange) ,and Nitrous hose line (Purple). Due

to the cost of nitrous, and budget constraints the testing of the feed system was limited.

However, it did reveal that temperature manipulation would be needed in order to obtain a

complete fill of the pressure vessel. Without the assistance of a pump or extreme

temperature manipulation it was deemed likely that only a ⅔ fill of the pressure vessel

would be obtainable and this was then shown to be true during actual testing.

6.2. Pressure Vessel Hydrostatic Test (Sean, Alex)

Prior to filling the pressure vessel with Nitrous Oxide, a Hydrostatic test needed to be

conducted to ensure the pressure vessel assembly could safely withstand the maximum

anticipated internal pressure of 1000 [psi]. To accomplish this, the pressure vessel inlet was

outfitted with a ¼ [in] NPT 1000 [psi] male to male ball valve, a T pipe fitting, and a

high-pressure gauge. Another 1000 [psi] ball valve was also added to the exit. The pressure
vessel was then taken to Industrial Service Solutions in Brighton Colorado where it was

pressurized with water to 1050 [psi] and observed for thirty minutes. During this time, no

visible leaking or reduction in internal pressure was detected. Figure 6.2.1 shows the

pressure vessel during this test.

Figure 6.2.1: Pressure Vessel Hydrostatic Test

After the successful completion of the Hydrostatic test, the pressure vessel was now able to

be filled with Nitrous Oxide.


6.3. Avionics (Garret, Alex)

6.3.1. Load Cell Calibration (Alex)

The load cell is an important part of the assembly as this is the component that will allow

the team to determine the thrust output of the system. With the load cell being oriented in a

new way compared to RocketLynx IV, the load cell must be calibrated to meet the new

conditions. Rather than the rocket system being in a vertical orientation where it is

clamped directly to the load cell, RocketLynx V has adjusted the system to be in a horizontal

position. With this, the load cell will no longer need to be calibrated to take the weight of

the combustion chamber into account.

To calibrate the load cell, the test stand was utilized as a clamping system for the load cell.

With the load cell clamped down, the team was able to utilize three known weights to

determine the readings of the load cell. The first weight that was placed on the load cell was

a total of 150 [lb] and with the load cell reading 70 [lb], the offsets in the code were

adjusted until the load cell was consistently reading 150 [lb]. The next weight used for

calibration was 120 [lb]. As we placed the 120 [lb] onto the load cell, the load cell was

consistently reading 120 [lb] which gave the team confidence. Finally, the final weight used

for calibration was 180 [lb]. With the load cell reading 180 [lb] consistently, the team was

confident to move forward with the load cell calibration.

With the load cell being calibrated, the team moved forward with confidence and mounted

the load cell to the new orientation of the test stand.


6.3.2. Pressure Transducers Calibration/Testing (Garrett)

The pressure transducers left over from RocketLynx 4 were in bad condition, so they were

replaced with new ones rated from 0 to 1000 [psi]. The pressure transducers were

successfully calibrated using the compressed shop air set to 100 [psi] with a pressure

regulator. The minimum voltage offset was adjusted in the code until the pressure reading

output from the Arduino code matched the pressure reading on the gauge connected to the

pressure regulator. While the Arduino was connected to a laptop via USB, the minimum

voltage was found to be 0.4 volts, which matched the specifications of the transducers. The

appendix contains the pressure transducer code.

After switching the Arduino power supply to a 9 volt battery, the transducer reading was

off. The avionics team found that changing the excitation voltage in the code to 3 volts,

down from 5 volts, fixed this calibration issue and the transducers were reading correctly

once again. Once the transducers were verified to work with the 9 volt battery, the team

was able to proceed with the nitrous flow test. This was done using a 10 [lb] nitrous bottle

directly connected to the run system. When the nitrous tank was opened, and the servos

were set to open, a pressure reading between 50 and 60 [psi] was observed.

The mass flow rate servo control also seemed to work throughout testing, however the run

system never reached the desired mass flow rate of 1 [kg/s] in any test, so the PID tuner set

the flow rate servo to remain fully open. This was true for all tests with the nitrous and

compressed air. If the avionics team set the desired mass flow rate to something smaller,

like 0.2 [kg/s], the servos would continually try to adjust to that mass flow rate, but they

would never actually settle on the desired number. Bernoulli’s equation for incompressible

flow was used to calculate the mass flow rates in all cases.
7. Hot-Fire Test Operations (Alex,Sean, Willis)

At the beginning of the Spring semester, the RocketLynx team was in contact with the

Colorado Air and Space Port (CASP) located in Watkins, Colorado, for all static system

testing. During conversations with the CASP, it was determined that prior agreements with

the RocketLynx teams could no longer continue, and new rules and regulations were put

into place for testing. Due to this, the team required an insurance plan to test the Hybrid

Rocket Propulsion System at CASP. The team sought insurance through the University of

Colorado Denver as this was a senior design project. However, this could not be obtained

due to the scale of the insurance required. Unfortunately, this news came to the team the

week before static testing occurred. Luckily, the team looked at multiple options for testing

and had contingency plans in place. With this, the static hot-fire test was scheduled at

Bandimere Speedway (Figure 7.1) in Morrison, Colorado.

Figure 7.1: Bandimere Speedway Location


The test was scheduled for Wednesday, May 3rd, at 9:00 am. The team first arrived at the

HUB located at the University of Colorado Denver to begin loading all equipment and tools

required for complete assembly. After this, the team arrived at the location and began the

setup process. Each team member had roles and responsibilities preassigned to ensure a

smooth setup process to limit the setup time between multiple burns. Once all setup was

complete, the Avionics team conducted a final check of all electrical components to ensure

the system would perform as expected, and data was successfully collected during the live

test.

Before the test began, the team closed all roads to the test location. The team members in

charge of ignition and the master on/off circuit waited for an all-clear before beginning

testing. As the all-clear was given, the test site was officially live for testing. As the

countdown was given, the ignition circuit was pressed, and 2 seconds later, the master

on/off was set to the on position. As the nitrous oxide began flowing into the combustion

chamber, ignition of the fuel grain was successfully attained, and the Hybrid Rocket

Propulsion system was on full blast (Figure 7.1). The entire burn lasted for 24 seconds and

ended with 36 seconds of bleeding, the pressure vessel empty. During the second test burn,

the pressure vessel was emptied for a safe refill procedure.


Figure 7.1: Burn Test Shortly After Ignition

During the disassembly process to begin the second test, the propulsion team opened up

the combustion chamber to switch out the ignition port. As the chamber was opened, it was

realized that the first test burn had completely burned through the fuel grain. With this

realization, all procedures were paused as a second burn test was no longer possible. Once

it was determined that a second test was no longer feasible, the teardown procedure was

initiated. The team began depressurization of the pressure vessel to ensure safe

transportation. Along with that, the team ensured no ignition sources were left in the

combustion chamber or testing location. Once these processes were achieved, the team

could pack up all materials and ensure the location was as clean as when we arrived.
8. Retrieved Data (Alex,Garrett) - Garrett add thrust

During testing, the data acquisition instrumentation successfully took data on the pressure

transducers, the mass flow rate, and the thrust output of the system. With the pressure

transducers being attached to the venturi device, pressure readings were taken to

determine the pressure of the nitrous oxide going through the run system as well as to

determine the mass flow rate. As can be seen in Figure 8.1, within the first 5 seconds of the

hot-fire static test, the pressure within the venturi device reaches a max of 335 [psi]. This

data shows that the master on/off circuit was working correctly as it was calibrated to open

the ball valve by 10 degrees at first. After a delay of 5 seconds, the ball valve moves into a

fully open position. Within the data, it is seen that as the ball valve opens to a fully open

position, the pressure within the run system begins to drop. At the ten second mark, the

pressure drops to 125 psi and as the flow of nitrous oxide continues for the 24 seconds of

burning, the pressure begins to drop back down to ambient.

Figure 8.1: Pressure Transducers Pressure Reading During Live Test


Using the pressure readings going through the venturi device, the mass flow rate was

calculated utilizing Bernoulli’s Principle. Looking at Figure 8.2, it can be seen that at the

beginning of the burn test, the mass flow rate spikes to 0.5 [kg/s] at the 2 second mark.

However, for the rest of the 24 second burn, the mass flow rate fluctuates between 0.1

[kg/s] to 0.3 [kg/s]. While the system was successful in adjusting the mass flow rate, it is

important to understand that the mass flow rate did not reach the desired flow for the

Hybrid Rocket Propulsion system.

Figure 8.2: Mass Flow Rate of Nitrous Oxide During Live Testing

In the pressure transducer section of the avionics, it was mentioned that the transducers

needed to be recalibrated after switching power supply to a 9 volt battery. This was not

done with the load cell before testing, so the raw data was skewed. The data acquisition

code listed in the appendix has the final and correct calibrations that should’ve been used

on test day. The team had to backtrack the test data to retrieve the original analog outputs

of each data point from testing. From there, the analog data was re-processed using the
correct calibration equations and coefficients to get the actual thrust output through the

load cell. Once this was done, it was found that the motor produced a peak thrust of 55.28

[lbf] at the very end of the burn, around 25 seconds. The data showed the thrust quickly

increased in the first 5 seconds of burn to 50 [lbf], and remained in the range of 45 to 54

[lbf] of thrust for the duration of the test.

Figure 8.3: Thrust Recorded from Load Cell During Live Testing

However, there could be some large error associated with this measurement. One issue

found was that the load cell continued to measure a steady amount of thrust after the motor

had burned away the fuel grain. This could be due to the test stand’s restraint cage

wrapping around the top of the combustion chamber being too tight. The team suspects

that the thrust pushed the combustion chamber back into the load cell, and the restraint

cage on the test stand kept it in that position until it was removed. This would also mean
the motor produced more thrust than the load cell recorded, but the combustion chamber

was being restrained by the test stand so it couldn’t move freely like it was intended to do.

Another source of error not accounted for was the friction between the bottom combustion

chamber and the test stand. A calculation was performed to estimate the thrust needed to

overcome the force of static friction due to the weight of the combustion chamber on the

test stand. It was found that approximately 18 [lbf] of friction force was exerted during the

live test fire. When accounting for this, the peak thrust of the rocket motor totals to 73.28

[lbf] of thrust.

9. Analysis of Data and Future Recommendations (Alex)

Following the hot-fire test operations, the team began to analyze the data as well as the

footage to determine the causes for failures during the testing of the system. During this

process, the team sought to identify what is known to work, the variables that affected

performance, and what subsystems dictate the control of those variables.

During the hot-fire test, the pressure vessel was successfully filled with nitrous oxide.

However, the pressure vessel was not filled to the desired amount of 15 [lb]. Along with

this, the expected pressure of 750 [lb] was not achieved within the pressure vessel. Instead,

the pressure vessel was filled with 10 [lb] of nitrous oxide and only reached a max pressure

of about 600 [psi]. While testing the feed system, this issue was seen but due to limits on

nitrous oxide, the team was not able to determine how detrimental this issue would be.

When analyzing the data provided by the pressure transducers and the mass flow rate

provided, it was seen that the mass flow rate of 1 [kg/s] was not achieved. Instead, the

control system was producing a mass flow rate between a range of 0.1 [kg/s] to 0.3 [kg/s]
for the 24 second burn time. The goal to reach 250 [lbf] of thrust relied heavily on being

able to achieve a mass flow rate of 1 [kg/s]. With the control system not being able to

achieve this, the thrust output of the system was under by a significant amount. Due to the

pressure vessel not filling to the desired amount as well as not reaching the desired

pressure, the team believes that the mass flow rate exiting the pressure vessel was greatly

under what calculations had shown us. With the design of the run system relying heavily on

gravity feed, pressure in the tank was vital to achieve an expected mass flow rate from the

system.

The propulsion team along with the avionics team was successful in being able to ignite the

fuel grain during the burn test. However, as the disassembly process began, it was realized

that the fuel grain burned completely away within one test fire. With this realization, the

team was not able to continue the burn test throughout the day. While the team had

originally calculated a burn time of 8 seconds, the test showed that the true burn time was

24 seconds. Within the 24 seconds burn time, the fuel grain was burned away entirely. By

analyzing the data provided by the avionics team as well as looking over the footage, the

team was able to determine that the pressure vessel was not emptying as quickly as we

expected. This was a result of the mass flow rate not reaching the desired amount of 1

[kg/s]. With the pressure vessel not emptying as quickly as expected, the fuel grain was

allowed to burn for the full 24 seconds of the burn test. Along with the fuel grain burning

up entirely, the insulating sleeve was burnt on the ends where the combustion began and

where the assembly transitioned into the nozzle. This ended up rendering the insulating

sleeve useless for future test runs.


Proven systems in play include the ignition circuit, ignition port, the master on/off, the

pressure vessel as a whole, the injector, the feed system, the run system, combustion

chamber, the nozzle, and finally the fuel grain. The ignition circuit was tested in a full scale

environment and successfully activated the electric initiator, which in turn, ignited the

ignition port. Along with this, the master on/off was used during the full scale burn test to

start and stop the flow of nitrous oxide into the combustion chamber. The feed system was

utilized to allow the team to fill the pressure vessel however, it did not fully fill to the

desired amount. The pressure vessel as a whole was hydrostatically tested and used during

the burn test as the source of nitrous oxide. The run system as a whole was successful in

transferring flow of the nitrous oxide into the injector and in turn into the combustion

chamber. With the run system successfully transferring nitrous oxide and the ignition port

being ignited properly, the fuel grain was also successful in reaching ignition once the

nitrous oxide was introduced into the process. The combustion chamber was successful in

achieving a combustion process while also withstanding all forces and temperatures

produced during the combustion process. Finally, the nozzle was successful in turning the

exhaust gas from subsonic to supersonic flow.

While most components functioned as intended, the team still understands that

improvements can be made to the current system before moving into future design

considerations. The first and best option would be to begin looking at new filling

techniques for the pressure vessel. While the feed system worked as intended, there were

still shortcomings that RocketLynx V could not overcome. As a team, we recommend adding

a pump for filling nitrous oxide fully into the pressure vessel. If a pump is not feasible,

utilizing dry ice as a cooling source for the tank will transfer nitrous oxide as it will flow
into the lower pressure system. The next improvement can be made within the control

system for the mass flow rate. By filling the pressure vessel to the desired amount, it is

possible to begin testing the control system away from the combustion process. This

presents an isolated change within the workflow of the motor, and is easily isolated so that

it can be monitored in the test environment. A design consideration that should be made is

utilizing a rocket phenolic liner rather than an insulating fiberglass sleeve. By making this

simple design change, it is possible to continue utilizing the liner for multiple tests rather

than having the insulating sleeve burn away during the testing process. Finally, RocketLynx

V recommends finding a new way to read the thrust output of the system. With the load cell

being mounted in front of the combustion chamber and directly onto the test stand, the

combustion chamber had to battle the forces of friction as well as the test stand cage

surrounding it. Adding sliding rails to the combustion chamber and possibly adding a steel

mounting plate for the load cell can allow for a more accurate thrust output reading.

With RocketLynx V redesigning, analyzing, and manufacturing an entirely new pressure

vessel, combustion chamber, nozzle, and fuel grain, the team found successes while

realizing areas where the system can improve. The team remains confident that a fully

realized system can be achieved during hot-fire testing, and understands that very few if

any first prototype tests work in their entirety. Going through this process not only allowed

the team to arrive at a workable solution to the current issue, but also identify potential

design improvements for future teams. It is important to note that multiple hypotheses for

the cause of failure were exploited, but the cause and solution listed above was identified as

the most likely, which is why it was pursued by the team.


10. Conclusion (Alex,Sean,Willis)

RocketLynx V began the year with the ambitious goals of designing, analyzing, and

manufacturing an all new hybrid rocket propulsion that could sustain 250 [lbf] of thrust for

8 seconds. While the team built off the developments from teams of years prior, major

fundamental design changes meant that most of the project started with a blank slate. Fall

semester started with redesigns and analysis of all major components. The Spring semester

included all manufacturing of all major components. With testing being completed towards

the end of the Spring semester, the team was able to achieve underlying goals which led to a

full 24 seconds burn test. While the RocketLynx V team was not able to reach all initial goals

set, the team established the viability of the Hybrid RocketLynx V hybrid rocket propulsion

system. With the recommendations given, the team believes that this system can be turned

into a fully capable engine ready for the Spaceport America Cup.

11. Acknowledgements

This project would not have been possible without the following individuals and

organizations. Junior members, David Malfavon Campos and Abraham Cabrera who worked

diligently on the ignition circuit along with Ruciel Hutapea who helped develop and

manufacture the test stand. Doug Gallagher who taught the team everything the team

needed to know in order to manufacture all components while also being an expert

consultant on the project. Bandimere Speedway for allowing the team to run a static test

fire at their location. Finally, all other students that helped by supporting the RocketLynx V

team within senior design during the year 2022-2023.


12. Budget

The RocketLynx V budget takes into account all funds for each senior team member and all

expenses used throughout the entire school year of 2022-2023. Tracking the budget for a

senior design team is an important metric for a project to succeed. Figure 12.1 shows that

almost all the budget for the RocketLynx V team was used for all stock materials for

manufacturing, all electronics, and all assembly materials for the system. The budget shown

is the final budget for the fifth iteration of RocketLynx.

Figure 12.1: RocketLynx V Budget


13. Gantt Chart

The Spring 2023 RocketLynx V Gantt chart (Figure 13.1) provides the ability for the team to

track progress on individual tasks, as well as monitor team progression as a whole.

Responsible team members are listed next to each item and this helped create

accountability within the team structure. The RocketLynx V Gantt chart has officially

reached the end as the project has come to an end.

Figure 13.1: RocketLynx V Gantt Chart


References

The American Society of Mechanical Engineers. (2017). Asme B16.5.

Bouziane, M., Bertoldi, A. E. M., Milova, P., Hendrick, P., & Lefebvre, M. (2019).

Performance comparison of oxidizer injectors in a 1-kn paraffin-fueled hybrid

rocket motor. Aerospace Science and Technology, 89, 392–406.

https://doi.org/10.1016/j.ast.2019.04.009

Chiaverini, M. J., & Kuo, K. K. (2007). Fundamentals of hybrid rocket combustion and

propulsion. American Institute of Aeronautics and Astronautics.

Combustion of composite propellants. (2015). Propellants and Explosives, 195–252.

https://doi.org/10.1002/9783527693481.ch7

Dennis, J., Hernandez, F., Shark, S., & Villarreal, J. (2010). Design of a N20/HTPB hybrid

rocket motor utilizing a toroidal aerospike nozzle. 48th AIAA Aerospace Sciences

Meeting Including the New Horizons Forum and Aerospace Exposition.

https://doi.org/10.2514/6.2010-183

Guo, H., Zhou, S., Shreka, M., & Feng, Y. (2019). Effect of pre-combustion chamber

nozzle parameters on the performance of a marine 2-stroke dual fuel engine.

Processes, 7(12), 876. https://doi.org/10.3390/pr7120876

Gupta, M. S., & Jha, E. (2018). Review on evaluation of properties of carbon phenolic

composite structure using gas chromatography for analysis of Thermal


performance. IOP Conference Series: Materials Science and Engineering, 455,

012041. https://doi.org/10.1088/1757-899x/455/1/012041

EDGE. (2018, March 3). RIT Multidisciplinary Engineering.

http://edge.rit.edu/edge/P18102/public/Engine%20Specifications/Propulsion/Combus

ti on%20Chamber/Combustion%20Chamber

McFarland, & Antunes. (2019). Small-scale static fire tests of 3D printing hybrid rocket

fuel grains produced from different materials. Aerospace, 6(7), 81.

https://doi.org/10.3390/aerospace6070081

Mechentel, F. S., & Cantwell, B. J. (2019). Experimental findings on pre- and

post-combustion chamber effects in a laboratory-scale motor. AIAA Propulsion

and Energy 2019 Forum. https://doi.org/10.2514/6.2019-4336

Nada, T. R., & Hashem, A. A. (2012). Geometrical characterization and performance

optimization of monopropellant thruster injector. The Egyptian Journal of

Remote Sensing and Space Science, 15(2), 161–169.

https://doi.org/10.1016/j.ejrs.2012.09.001

Nagata, H., Hagiwara, S., Wakita, N., Totani, T., & Uematsu, T. (2011). Optimal Fuel

Grain design method for Camui Type Hybrid Rocket. 47th AIAA/ASME/SAE/ASEE

Joint Propulsion Conference &Amp; Exhibit.

https://doi.org/10.2514/6.2011-6105
Ozawa, K., & Shimada, T. (2020). Performance of mixture-ratio-controlled hybrid

rockets for nominal fuel regression. Journal of Propulsion and Power, 36(3),

400–414. https://doi.org/10.2514/1.b37665

Pastor, T. (2013). Section VIII—division 1: Rules for construction of pressure vessels.

Companion Guide to the ASME Boiler & Pressure Vessel Code.

https://doi.org/10.1115/1.859872.ch21

Qazi, M. U., Masood, K., & Mahmood, K. (2009). Rapid Prediction of Star Propellant

Grain Regression using CAD Software. International Bhurban Conference on

Applied Sciences & Technology, 244–248.

Redwood, B., Schöffer Filemon, Garrett, B., & Fadell, T. (2020). The 3D Printing

Handbook Technologies, design and applications. 3D Hubs.

Sutton, G. P., & Biblarz, O. (2017). Rocket Propulsion Elements (9th ed.). Wiley.

Taylor Forge, G+w Taylor-Bonney Division. (n.d.). Modern Flange Design.

Tian, H., Li, X., & Cai, G. (2015). Ignition theory investigation and experimental

research on Hybrid Rocket Motor. Aerospace Science and Technology, 42,

334–341. https://doi.org/10.1016/j.ast.2015.01.015

Whitmore, S. A., Sobbi, M., & Walker, S. (2014). High regression rate hybrid rocket fuel

grains with helical port structures. 50th AIAA/ASME/SAE/ASEE Joint Propulsion

Conference. https://doi.org/10.2514/6.2014-3751
Whitmore, S. A. (2020). Nytrox as “drop-in” replacement for gaseous oxygen in

SmallSat hybrid propulsion systems. Aerospace, 7(4), 43.

https://doi.org/10.3390/aerospace7040043

Whitmore, S., Peterson, Z., & Eilers, S. (2011). Analytical and experimental

comparisons of HTPB and ABS as hybrid rocket fuels. 47th

AIAA/ASME/SAE/ASEE Joint Propulsion Conference &Amp; Exhibit.

https://doi.org/10.2514/6.2011-5909

Zandbergen, B. T. C. (2010). Thermal Rocket Propulsion (version 2.04): AE4S01. TU

Delft.

Zhu, H., Li, M., Tian, H., Wang, P., Yu, N., & Cai, G. (2019). Numerical and experimental

investigations on injection effects of orifice injector plate in hybrid rocket motors. Acta

Astronautica, 162, 275–283. https://doi.org/10.1016/j.actaastro.2019.06.002


Appendix

Data Acquisition Code

// DATA ACQUISITION CODE -- Final Version w/ corrected calibrations


// THIS ARDUINO CONNECTED TO LAPTOP
// ALL OTHERS ON 9 VOLT POWER SUPPLY

// transducers
float voltage1;
float voltage2;

// TRANSDUCER CALIBRATIONS
float transducerMinVolts = 0.4; // volts, spec sheet CALIBRATED
//float transducerMinVolts = 0.69; // volts, spec sheet
float transducerMaxVolts = 4.5; // volts, spec sheet
float analogRange = 1024.0; // max analog range
//float inputVolts = 5.0; // volts, coming from arduino LAPTOP CALIBRATION
float inputVolts = 3.0; // volts, coming from arduino TEST DAY CALIBRATION
float transducerMinAnalog = analogRange*(transducerMinVolts/inputVolts);
float transducerMaxAnalog = analogRange*(transducerMaxVolts/inputVolts);
float transducerMinRange = 0; //psi
float transducerMaxRange = 1000; //psi

//float density = 0.00029489; // lb/(in^3), shop air at 100 psi, room temp
float density = 0.000053611; // lb/(in^3), NOS
// select densities above based on what fluid is being used to get correct mass
flow rate

float meter2inchConversion = 39.3701; //inch/meter


float D1meters= .0178594; //m VENTURI
float D1 = D1meters*meter2inchConversion; // inches
float D2meters= .009525; //m VENTURI
float D2 = D2meters*meter2inchConversion; // inches
float beta = pow(D2/D1, 4); // unitless ratio, diameters 2/1

float lb2kgConversion = 1.0/2.2046244201838;


// LOAD CELL CALIBRATIONS
float a0 = 0.037142857;
float a1 = 0.002053061;
float K = 1/a1;
// float ex = 5; // EXCITATION VOLTAGE, OLD CALIBRATION
float ex = 3; // EXCITATION VOLTAGE, NEW CALIBRATION CORRECTED FOR TEST DAY
const float referenceVolts = 9.0;

void setup() {
Serial.begin(9600);
float A1 = (PI/4.0)*pow(D1, 2); // inches^2
float A2 = (PI/4.0)*pow(D2, 2); // inches^2
delay(1000);
}
float counter=1;
float sum=0;
float avg=0;

void loop() {
Serial.println("--------------------------------");

// TRANSDUCERS AND MASS FLOW RATE


float dPressure = Pressure();

float massFlowRate = 0.0;


float massFlowRateSI = 0.0;
massFlowRate = A2*sqrt( (2.0*density*dPressure)/(1.0-beta) ); // lb/s
massFlowRateSI = lb2kgConversion*massFlowRate; // kg/s
//Serial.print("Pressure Change (psi): "); Serial.println(dPressure, 2);
//Serial.print("Mass Flow Rate: "); Serial.println(massFlowRate, 2);
Serial.print("Mass Flow Rate SI (kg/s): "); Serial.println(massFlowRateSI, 2);

// LOAD CELL
float val1 = analogRead(A2); // read the value from the sensor
//Serial.println(val1);
float volts = (val1 / 1023.0) * referenceVolts; // calculate the ratio
float yi = volts/ex;
float weight = ((K*yi)-(K*a0));
sum+=weight;
if(counter==2){
avg=(sum/(counter+1))-22; //-8 is offset. Set different if needed
counter = 0;
sum = 0;
//Serial.print("loadcell ");Serial.println(avg);
}
Serial.print("loadcell: ");Serial.println(avg);
//Serial.print("Counter ");Serial.println(counter);
counter ++;
delay(100);
}

float Pressure(){

float PressureTrans1= analogRead(A0);


float PressureTrans2= analogRead(A1);

voltage1= PressureTrans1*(inputVolts/analogRange);
voltage2= PressureTrans2*(inputVolts/analogRange);

float pressure1 = transducerMaxRange*(PressureTrans1 -


transducerMinAnalog)/(transducerMaxAnalog-transducerMinAnalog);
float pressure2 = transducerMaxRange*(PressureTrans2 -
transducerMinAnalog)/(transducerMaxAnalog-transducerMinAnalog);

Serial.print("Pressure 1 (psi): "); Serial.println(pressure1, 2);


Serial.print("Pressure 2 (psi): "); Serial.println(pressure2, 2);

float PressureChange = abs(pressure1-pressure2);


//Serial.print("Transducer 1 Analog: "); Serial.println(PressureTrans1);
//Serial.print("Transducer 2 Analog: "); Serial.println(PressureTrans2);
//Serial.print("Voltage 1: "); Serial.println(voltage1);
//Serial.print("Voltage 2: "); Serial.println(voltage2);

return PressureChange;
}
Pressure Transducer Code with PID Servo Control for Nitrous

// PRESSURE TRANSDUCER CODE FOR NITROUS W SERVO PID CONTROL


// 9 VOLT BATTERY ARDUINO POWER SUPPLY

// Servo
#include<Servo.h>
#include<Wire.h>
Servo BallServo;
#define ServoPin 3
float desireFlow = 1.0; //kg/sec
float desireError = 0.0;
float FlowError; // desire value - actual value
float ServoPID;
float ki=-70; // integral controller
float kp=500; // proportional controller
float kd=300; // derivative controller
float ServoInput; // voltage needed to go in servo to regulate flow in pipe
float PID_ki,PID_kd, PID_kp;
float dt; //change in time
float OldError;
unsigned long NewTime,Oldtime;
float ErrorArea=0;
float PID;
float error=0;

// transducers
float voltage1;
float voltage2;
float transducerMinVolts = 0.4; // volts, spec sheet
//float transducerMinVolts = 0.24; // volts, spec sheet
float transducerMaxVolts = 4.5; // volts, spec sheet
float analogRange = 1024.0; // max analog range
float inputVolts = 5.0; // volts, coming from arduino
float transducerMinAnalog = analogRange*(transducerMinVolts/inputVolts);
float transducerMaxAnalog = analogRange*(transducerMaxVolts/inputVolts);
float transducerMinRange = 0; //psi
float transducerMaxRange = 1000; //psi
//float density = 0.00029489; // lb/(in^3), shop air at 100 psi, room temp
float density = 0.000053611; // lb/(in^3), NOS

float meter2inchConversion = 39.3701; //inch/meter


float D1meters= .0178594; //m VENTURI
float D1 = D1meters*meter2inchConversion; // inches
float D2meters= .009525; //m VENTURI
float D2 = D2meters*meter2inchConversion; // inches
float beta = pow(D2/D1, 4); // unitless ratio, diameters 2/1

float lb2kgConversion = 1.0/2.2046244201838;

void setup() {
Serial.begin(9600);
float A1 = (PI/4.0)*pow(D1, 2); // inches^2
float A2 = (PI/4.0)*pow(D2, 2); // inches^2
BallServo.attach(ServoPin);
BallServo.write(90);
delay(1000);
NewTime= millis();
}

void loop() {
Serial.println("--------------------------------");

float dPressure = Pressure();


float massFlowRate = 0.0;
float massFlowRateSI = 0.0;
massFlowRate = A2*sqrt( (2.0*density*dPressure)/(1.0-beta) ); // lb/s
massFlowRateSI = lb2kgConversion*massFlowRate; // kg/s

Serial.print("Pressure Change (psi): "); Serial.println(dPressure, 2);


//Serial.print("Mass Flow Rate: "); Serial.println(massFlowRate, 2); // FOR
ENGLISH
Serial.print("Mass Flow Rate SI (kg/s): "); Serial.println(massFlowRateSI, 2);

PIDContollers(PID_kp, PID_ki,PID_kd, massFlowRateSI);


PID= abs((kp*error + kd*PID_kd + ki*PID_ki));
if (massFlowRateSI < desireFlow){
PID=90;
}
BallServo.write(PID);
Serial.print("PID: ");Serial.println(PID);

delay(1000);
}

float PIDContollers(float P, float I, float D, float CurrentFlow){

error=abs( desireFlow-CurrentFlow);
Serial.print("error: ");Serial.println(error);
Oldtime= NewTime;
NewTime=millis();
dt= NewTime - Oldtime;
error= desireFlow-CurrentFlow;
D =(error-OldError)/dt;
ErrorArea= ErrorArea + error*dt;
OldError= error;
}

float Pressure(){

float PressureTrans1= analogRead(A0);


float PressureTrans2= analogRead(A1);
voltage1= PressureTrans1*(inputVolts/analogRange);
voltage2= PressureTrans2*(inputVolts/analogRange);

float pressure1 = transducerMaxRange*(PressureTrans1 -


transducerMinAnalog)/(transducerMaxAnalog-transducerMinAnalog);
float pressure2 = transducerMaxRange*(PressureTrans2 -
transducerMinAnalog)/(transducerMaxAnalog-transducerMinAnalog);
Serial.print("Pressure 1 (psi): "); Serial.println(pressure1, 2);
Serial.print("Pressure 2 (psi): "); Serial.println(pressure2, 2);
float PressureChange = abs(pressure1-pressure2);
Serial.print("Transducer 1 Analog: "); Serial.println(PressureTrans1);
Serial.print("Transducer 2 Analog: "); Serial.println(PressureTrans2);
Serial.print("Voltage 1: "); Serial.println(voltage1);
Serial.print("Voltage 2: "); Serial.println(voltage2);

return PressureChange;
}

Master On/Off Code

#include <Servo.h>

Servo myservo; // create servo object to control a servo

// twelve servo objects can be created on most boards

int pos = 0; // variable to store the servo position

void setup() {

myservo.attach(10); // attaches the servo on pin 9 to the servo object

//for (int i = 65; i >= 0; i--) {

// myservo.write(i);

// delay(10);

//}

pinMode(6, INPUT);
pinMode(7, INPUT);

pinMode(8,INPUT);

pinMode(9,INPUT);

int keep = 2 ;

void loop() {

if (digitalRead(7) == LOW) {

keep = 1;

pos=0;

if (digitalRead(9) == LOW) {

keep = 0;

pos=100;

// for (pos = 0; pos <= 180; pos += 1) { // goes from 0 degrees to 180 degrees

// in steps of 1 degree
while (keep == 1) {

//if(pos == 90) {

// delay(1000);

// }

digitalWrite(6,HIGH);

digitalWrite(8,LOW);

pos+=5;

myservo.write(pos);

Serial.print("ON: "); Serial.println(pos); // tell servo to go to position in variable 'pos'

delay(10);

if (pos == 100) {

keep = 2;

while (keep == 0) {

if(pos == 80) {

delay(2000);

digitalWrite(6,LOW);

digitalWrite(8,HIGH);

pos-=5;
myservo.write(pos);

Serial.print("OFF: "); Serial.println(pos); // tell servo to go to position in variable 'pos'

delay(10);

if (pos == 0) {

keep = 2;

You might also like