[go: up one dir, main page]

0% found this document useful (0 votes)
53 views89 pages

Project Report

This project report focuses on optimizing the cryogenic treatment and aging parameters for 17-4 PH stainless steel to enhance its mechanical properties and corrosion resistance. The study evaluates various heat treatment processes, including standard aging, double aging, and cryogenic aging, revealing that different treatments yield specific benefits in strength, ductility, and toughness. The findings aim to tailor heat treatment methods for application-specific requirements in engineering fields such as aerospace and petrochemical industries.

Uploaded by

Jai Aggarwal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
53 views89 pages

Project Report

This project report focuses on optimizing the cryogenic treatment and aging parameters for 17-4 PH stainless steel to enhance its mechanical properties and corrosion resistance. The study evaluates various heat treatment processes, including standard aging, double aging, and cryogenic aging, revealing that different treatments yield specific benefits in strength, ductility, and toughness. The findings aim to tailor heat treatment methods for application-specific requirements in engineering fields such as aerospace and petrochemical industries.

Uploaded by

Jai Aggarwal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 89

OPTIMIZATION OF CRYOGENIC TREATMENT AND AGING

PARAMETERS FOR 17-4 PH STAINLESS STEEL

A PROJECT REPORT

Submitted by
AMAN KADU (112111005)
SALONI SAKALA (112111052)

In the fulfillment for the award of the degree of

BACHELOR OF TECHNOLOGY

IN

METALLURGY AND MATERIAL TECHNOLOGY

COLLEGE OF ENGINEERING, PUNE 411005

(MAY 2025)

i
DEPARTMENT OF METALLURGY
AND MATERIAL ENGINEERING
COLLEGE OF ENGINEERING, PUNE 411005 (INDIA)

CERTIFICATE

This is certified that the project report “OPTIMIZATION OF CRYOGENIC


TREATMENT AND AGING PARAMETERS FOR 17-4 PH STAINLESS STEEL” is the
bonafide work of “AMAN KADU and SALONI SAKALA” who carried out the project work
under my guidance.

SIGNATURE SIGNATURE

Dr. M. G. Kulthe. Dr. S. U. Dangrikar

HEAD OF DEPARTMENT PROJECT GUIDE

Metallurgy and Material Engineering, Metallurgy and Material Engineering,


College of Engineering, Pune-411005 College of Engineering, Pune-411005

ii
Contents
TITLE PAGE ...................................................................................................................................... i

CERTIFICATE ...................................................................................................................................... ii

ABSTRACT......................................................................................................................................... vii

SYMBOLS.......................................................................................................................................... viii

CHAPTER 1 .......................................................................................................................................... 1

INTRODUCTION ................................................................................................................................. 1

1.1 Overview of 17-4 PH Stainless Steel ...................................................................................... 1

1.2 Background of 17-4 PH SS ..................................................................................................... 2

1.3 History of 17-4 PH SS............................................................................................................. 2

1.4 Current Status/ Applications ................................................................................................... 3

1.5 Organization of Work.............................................................................................................. 4

CHAPTER 2 .......................................................................................................................................... 5

LITERATURE REVIEW ...................................................................................................................... 5

2.1 Optimizing Heat Treatment Parameters for 17-4 PH Stainless Steel...................................... 5

2.1.1 Importance of Stainless Steels in Industry ....................................................................... 5

2.1.2 Precipitation-Hardening Stainless Steels (PHSS) ............................................................ 5

2.1.3 17-4 PH Stainless Steel: A Widely Used PH Grade ........................................................ 6

2.1.4 Objective of the Study ..................................................................................................... 6

2.1.5 Scope of the Literature Review........................................................................................ 6

2.2 Metallurgy of 17-4 PH Stainless Steel .................................................................................... 7

2.2.1 Chemical Composition and Alloying Elements ............................................................... 7

2.2.2 Phase Transformations ..................................................................................................... 9

2.2.3 Microstructural Constituents .......................................................................................... 10

2.3 Conventional Heat Treatment of 17-4 PH Stainless Steel .................................................... 11

2.3.1 Solution Annealing ........................................................................................................ 11


iii
2.3.2 Standard Aging .............................................................................................................. 12

2.4 Advanced Heat Treatment Techniques for 17‑4 PH Stainless Steel ..................................... 14

2.5 Influence of Heat Treatments on 17-4 PH Stainless Steel .................................................... 18

2.5.1 Influence on Microstructure ........................................................................................... 18

2.5.2 Influence on Mechanical Properties............................................................................... 19

2.5.3 Influence on Corrosion Resistance ................................................................................ 20

2.5.4 Influence on Fracture Behavior ..................................................................................... 21

2.6 Characterization Techniques for Heat-Treated 17‑4 PH Stainless Steel............................... 22

2.6.1 Tensile Testing ............................................................................................................... 22

2.6.2 Impact Testing (Charpy) ................................................................................................ 23

2.6.3 Hardness Testing ............................................................................................................ 23

2.6.4 Microstructural Analysis ................................................................................................ 24

2.6.5 Fracture Analysis (SEM) ............................................................................................... 25

2.6.6 Corrosion Testing........................................................................................................... 26

2.7 Research Gaps and Motivation ............................................................................................. 27

2.8 Conclusion of Literature Review .......................................................................................... 28

2.9 Experimental Flowchart ........................................................................................................ 29

CHAPTER 3 ........................................................................................................................................ 30

EXPERIMENTAL METHODOLOGY ............................................................................................... 30

3.1 Material Procurement ............................................................................................................ 30

3.2 Chemical Analysis of Sample ............................................................................................... 30

3.3 Heat Treatment Cycles .......................................................................................................... 31

3.4 Microstructural Evolution ..................................................................................................... 35

3.5 Rockwell Hardness Testing ................................................................................................... 37

3.6 Microhardness Testing .......................................................................................................... 39

3.7 Testing of Charpy Impact Energy ......................................................................................... 40

3.8 Determination of YS, UTS, %Elongation ............................................................................. 42

.............................................................................................................................................................. 43
iv
3.9 X-Ray Diffractometer Testing .............................................................................................. 44

3.10 Corrosion Resistance Testing (Tafel Extrapolation) ............................................................. 45

CHAPTER 4 ........................................................................................................................................ 47

RESULTS AND DISCUSSION .......................................................................................................... 47

4.1 Microstructure Analysis ........................................................................................................ 47

4.1.1 AR Sample ..................................................................................................................... 47

4.1.2 SA Sample ..................................................................................................................... 48

4.1.3 DA1B Sample ................................................................................................................ 48

4.1.4 DA2B Sample ................................................................................................................ 49

4.1.5 CA1B Sample ................................................................................................................ 49

4.1.6 CA2B Sample ................................................................................................................ 50

4.1.7 CA3B Sample ................................................................................................................ 50

4.2 Rockwell Hardness Results ................................................................................................... 51

4.3 Microhardness Result Analysis ............................................................................................. 51

4.4 Charpy Impact Energy Analysis ........................................................................................... 52

4.5 Analysis Of YS, UTS and %Elongation ............................................................................... 53

4.5.1 Yield Strength ................................................................................................................ 53

4.5.2 Ultimate Tensile Strength .............................................................................................. 54

4.5.3 %Elongation ................................................................................................................... 55

4.5.4 Stress-Strain Curve ........................................................................................................ 55

4.6 XRD Analysis ....................................................................................................................... 57

4.7 Corrosion Resistance Analysis .............................................................................................. 67

CHAPTER 5 ........................................................................................................................................ 69

CONCLUSION .................................................................................................................................... 69

5.1 Microstructural Findings:............................................................................................... 69

5.2 Mechanical Property Summary: .................................................................................... 69

5.3 XRD and Phase Analysis: .............................................................................................. 70

5.4 Corrosion Resistance (Tafel Test): ................................................................................ 70


v
5.5 Overall Conclusion: ....................................................................................................... 70

5.6 Final Verdict: ................................................................................................................. 70

CHAPTER 6 ........................................................................................................................................ 72

FUTURE SCOPE................................................................................................................................. 72

APPENDICES ..................................................................................................................................... 73

Appendix A: Hardness Of 17-4 PH SS ............................................................................................ 73

Appendix B: Microhardness of 17-4 PH SS .................................................................................... 74

Appendix C: YS, UTS, %Elongation Of 17-4 PH SS...................................................................... 75

Appendix D: Charpy Impact Test Of 17-4 PH SS ........................................................................... 76

REFRENCES ....................................................................................................................................... 78

ACKNOWLEDGEMENT ................................................................................................................... 81

vi
ABSTRACT

17-4 PH stainless steel is a widely used precipitation-hardening alloy known for its excellent combination of
strength, corrosion resistance, and toughness. However, its mechanical and chemical performance is highly
sensitive to heat treatment conditions. This project aimed to optimize the heat treatment parameters of 17-4
PH stainless steel by evaluating the effects of standard aging, double aging, and cryogenic aging processes on
its microstructure, mechanical properties, phase constitution, and corrosion resistance.

The experimental work involved solution annealing at 1040 °C for 30 minutes followed by different secondary
treatments: standard aging at 480 °C for 1 hour; double aging sequences at 480 °C/2 h + 550–570 °C/4 h; and
cryogenic treatments using −196 °C liquid nitrogen for varying durations (3–9 h) prior to aging. A total of
seven heat treatment batches were prepared. Samples were subjected to mechanical testing (tensile strength,
yield strength, hardness, impact toughness, elongation), microstructural characterization (optical microscopy,
SEM fracture analysis), phase identification (XRD), and electrochemical corrosion testing (Tafel
extrapolation).

Results showed that standard aging (SA) produced the highest yield strength (0.812 kN) due to the formation
of fine, coherent Cu-rich precipitates. Double aging (DA2B) led to coarser lath structures and reduced hardness
but provided the best ductility (19.8%) and impact toughness (60.97 J), attributed to effective martensitic
tempering. Cryogenic aging (especially CA2B) refined the martensite and produced the highest hardness
(48.66 HRC) and tensile strength (1.93 kN), although with moderate corrosion resistance. XRD confirmed full
martensitic transformation in all heat-treated samples, while Tafel analysis revealed that the as-received (AR)
condition had the best corrosion resistance (0.145 mm/year), and aggressive treatments like CA3B and double
aging reduced corrosion performance.

The study concludes that heat treatment can be tailored to application-specific requirements: CA2B for high-
strength components, DA2B for impact-resistance, and SA for general structural use. This work highlights the
critical role of thermal and cryogenic parameters in optimizing the structure–property–corrosion relationships
in 17-4 PH stainless steel.

vii
SYMBOLS

Sr. no. Symbols Nomenclature

1. AR As Received

2. SA Standard Aging

3. DA1B Double Aging 1 Batch

4. DA2B Double Aging 2 Batch

5. CA1B Cryogenic Aging 1 Batch

6. CA2B Cryogenic Aging 2 Batch

7. CA3B Cryogenic Aging 3 Batch

viii
CHAPTER 1
INTRODUCTION

17-4 precipitation-hardening (PH) stainless steel is a martensitic grade known for its exceptional combination
of high strength, hardness, and corrosion resistance, making it a widely used material in aerospace, nuclear,
petrochemical, and marine applications. Its mechanical properties can be tailored through controlled heat
treatment processes, particularly solution treatment followed by aging.

The present study focuses on samples of 17-4 PH stainless steel subjected to cryogenic treatment and aging
cycle optimization to enhance performance. Historically developed in the mid-20th century to meet the
demands of high-strength and corrosion-resistant applications, 17-4 PH has seen continuous evolution in its
processing techniques. Today, it remains a material of choice in critical applications requiring dimensional
stability and wear resistance. This report explores the metallurgical background, processing history, and
current industrial relevance of 17-4 PH, laying the foundation for experimental analysis and optimization in
the chapters that follow.

1.1 Overview of 17-4 PH Stainless Steel

17-4 precipitation-hardening (PH) stainless steel, also known as AISI 630 or UNS S17400, is a martensitic
stainless-steel alloy widely used for its exceptional combination of high strength, hardness, and corrosion
resistance. The term "17-4" indicates its nominal composition of approximately 17% chromium and 4% nickel,
along with other critical alloying elements such as copper (Cu), manganese (Mn), silicon (Si), and niobium
(Nb), also known as columbium. These alloying elements contribute to its excellent precipitation-hardening
response and overall mechanical performance.

Unlike traditional austenitic or ferritic stainless steels, 17-4 PH undergoes a unique sequence of thermal
treatments that include solution annealing followed by aging to induce precipitation of intermetallic phases,
primarily copper-rich particles. This precipitation hardening transforms the soft, ductile martensite into a much
harder and stronger phase without significantly sacrificing its corrosion resistance, making it suitable for a
variety of demanding service environments.

Another notable feature of this alloy is its ease of fabrication and heat treatment. It can be machined in the
annealed (solution-treated) condition and then strengthened to various aging conditions (e.g., H900, H1025,
H1150) depending on the application’s requirements. Each condition varies in aging temperature and time,
resulting in different balances of strength, hardness, and ductility.

1
The microstructure of 17-4 PH stainless steel is predominantly martensitic after solution treatment, and the
subsequent aging results in fine precipitates which enhance the strength and stability of the alloy. In addition,
this alloy exhibits good corrosion resistance in both oxidizing and reducing environments due to its chromium
content. It resists stress corrosion cracking better than many austenitic grades, especially in chloride-containing
environments.

1.2 Background of 17-4 PH SS

This study utilizes commercially available 17-4 PH stainless steel in its solution-annealed condition (Condition
A), which offers a consistent martensitic microstructure suitable for further thermal and cryogenic processing.
Known for its strength and corrosion resistance, 17-4 PH becomes precipitation-hardened upon aging due to
the formation of copper-rich intermetallics. The material's initial state allows for clear observation of
microstructural and mechanical changes induced by subsequent treatments. Cryogenic treatment, typically
performed at −150 °C to −196 °C using liquid nitrogen, is integrated into the heat treatment cycle to transform
retained austenite to martensite and refine the microstructure, with the goal of improving properties like
hardness, wear resistance, and fatigue life.

This project investigates the combined influence of cryogenic treatment and varying aging parameters on the
alloy’s mechanical behaviour and microstructure. Aging temperatures and durations are systematically varied
both with and without prior cryogenic exposure. Properties such as tensile strength, hardness, impact toughness
are evaluated. Microstructural changes are studied through SEM imaging and phase analysis using XRD. This
approach aims to optimize the thermal treatment cycle to enhance the overall performance of 17-4 PH stainless
steel for advanced engineering applications.

1.3 History of 17-4 PH SS

The development of 17-4 PH stainless steel dates back to the 1940s, during efforts to design high-performance
alloys for military and aerospace applications. At that time, there was a need for a material that could offer the
high strength of martensitic steels and the corrosion resistance of austenitic grades—properties not typically
found together in conventional stainless steels. Developed by Armco Steel (now AK Steel), 17-4 PH was
designed around a martensitic matrix with copper-based precipitation hardening, which allowed for significant
enhancement of mechanical properties through controlled heat treatment. The inclusion of niobium helped
refine grain structure and improve thermal stability, making the alloy easier to process and more versatile.

2
Over the decades, the alloy gained widespread industrial acceptance due to its adaptability and cost-
effectiveness. By the mid-20th century, it had become a standard material in the aerospace, petrochemical, and
nuclear sectors. Research and process improvements continued through the late 20th and early 21st centuries,
focusing on optimizing aging cycles and exploring advanced treatments like double aging and cryogenic
processing. These refinements have allowed engineers to tailor the alloy’s performance to meet increasingly
demanding operational conditions, ensuring that 17-4 PH remains a critical material in both legacy systems
and emerging technologies.

1.4 Current Status/ Applications

17-4 PH stainless steel remains one of the most widely used precipitation-hardening grades in engineering
applications. Its performance characteristics make it highly attractive in fields that require excellent
mechanical strength, corrosion resistance, and good weldability. Notably, this material is used in the aerospace
sector for manufacturing turbine engine parts, structural fittings, shafts, and fasteners that are exposed to high
mechanical and thermal stress. Its ability to retain strength and dimensional accuracy at elevated temperatures
makes it suitable for these critical applications.

In the nuclear and marine industries, 17-4 PH is used for pump shafts, valves, nuclear reactor internals, and
other components that are exposed to corrosive environments and radiation. The medical field also benefits
from this alloy's biocompatibility and sterilization resistance, using it in surgical instruments, orthopaedic
implants, and dental devices. Oil and gas industries utilize it in downhole tools, pressure vessels, and other
components that require a high strength-to-corrosion resistance ratio.

In recent years, there has been growing interest in further enhancing the performance of 17-4 PH through
advanced heat treatment techniques. Among these, cryogenic treatment has emerged as a promising post-
processing step capable of improving wear resistance, fatigue life, and dimensional stability. The integration
of cryogenic processing with optimized aging cycles can result in microstructural refinements that enhance
mechanical performance beyond what is typically achievable through standard heat treatment alone.

This research project seeks to explore this frontier by evaluating how varying aging parameters in combination
with cryogenic treatment influence the behaviour of 17-4 PH stainless steel. Through mechanical testing and
microstructural characterization, the study aims to develop a deeper understanding of phase transformations,
precipitation behaviour, and the resulting mechanical performance. The results could have practical
implications for industries where even marginal improvements in component reliability and service life are of
critical importance.

3
1.5 Organization of Work

Chapter 1: Introduction – Provides an overview of 17-4 PH stainless steel, its composition, properties, and
relevance to modern engineering applications. It also outlines the background of the study, historical
development of the alloy, and current industrial significance. The motivation and objectives of the project are
presented.

Chapter 2: Literature Review – Summarizes existing research on heat treatment and cryogenic processing
of 17-4 PH stainless steel. It highlights the effects of various aging and cryogenic parameters on mechanical
properties and microstructure, identifying gaps that this study aims to address.

Chapter 3: Experimental Methodology – Describes the preparation and treatment of samples, detailing the
cryogenic treatment cycle, aging schedules, and testing methodologies. Equipment used and standard
procedures for tensile testing, hardness measurement, wear testing, and fatigue analysis are outlined.

Chapter 4: Results and Discussion – Presents the mechanical test results, along with SEM and XRD
microstructural analysis. The influence of different aging and cryogenic conditions on material performance
is discussed in detail, supported by comparative analysis and graphical representations.

Chapter 5: Conclusions – Summarizes the key findings and draws conclusions regarding the effect of
cryogenic treatment and optimized aging on the mechanical and structural behaviour of 17-4 PH stainless
steel.

Chapter 6: Future Work – Suggests future directions for research, including extended cryogenic holding
times, alternative aging profiles, and exploration of other precipitation-hardenable alloys under similar
treatment schemes.

4
CHAPTER 2
LITERATURE REVIEW
2.1 Optimizing Heat Treatment Parameters for 17-4 PH Stainless Steel

2.1.1 Importance of Stainless Steels in Industry

Stainless steels are iron-based alloys containing a minimum of 10.5% chromium, which forms a passive oxide
layer that imparts excellent corrosion resistance [1]. Owing to their outstanding mechanical strength, resistance
to oxidation and corrosion, and versatility in fabrication, stainless steels find applications across a broad
spectrum of industries. These include:

• Aerospace: components like landing gear, actuators, and engine parts.


• Medical: surgical instruments, orthopedic implants due to biocompatibility.
• Chemical processing: reactors, heat exchangers, pipelines for corrosive environments.
• Food and beverage industry: due to hygienic and non-reactive surfaces.
Among the different types of stainless steels—ferritic, austenitic, martensitic, duplex, and precipitation
hardening (PH)—the PH category is distinct due to its ability to achieve high strength through aging treatments
at relatively lower temperatures [2].

2.1.2 Precipitation-Hardening Stainless Steels (PHSS)

PH stainless steels combine properties of martensitic or semi-austenitic matrices with strengthening via
precipitation of fine secondary phases, typically Cu-rich or Ni-Al particles, during aging treatments [3]. Their
benefits include:

• High tensile and yield strength


• Good corrosion resistance (better than martensitic steels)
• Dimensional stability
• Ease of machinability in the solution-treated state
These attributes make them attractive in critical applications like aerospace structural components, nuclear
fuel reprocessing, and high-pressure valve parts [4].

PHSS are generally classified into:

• Martensitic PH steels (e.g., 17-4 PH, 15-5 PH): Higher strength, used where load-bearing capability
is critical.
• Austenitic PH steels (e.g., A-286): Better formability and toughness.

5
2.1.3 17-4 PH Stainless Steel: A Widely Used PH Grade

17-4 PH (UNS S17400) is the most widely used PH stainless steel. It contains approximately 17% chromium
and 4% nickel, with additions of copper and niobium for precipitation strengthening. Its microstructure
typically transforms to martensite upon solution treatment followed by air cooling. Aging treatments then
result in fine Cu-rich precipitates that enhance strength.

Key properties:

• Tensile Strength: 930–1310 MPa


• Yield Strength: 800–1180 MPa
• Hardness: up to 44 HRC
• Corrosion resistance: comparable to austenitic grades in many environments [5]
Its property balance makes it suitable for:

• Aircraft fittings
• Nuclear waste casks
• Marine shafts
• Surgical tools
However, the performance of 17-4 PH steel is highly sensitive to heat treatment parameters. Variations in
aging temperature, time, and sequence (e.g., double aging or cryogenic treatments) can significantly affect the
size, distribution, and nature of precipitates, thereby altering mechanical and corrosion performance [6].

2.1.4 Objective of the Study

The primary objective of this project is to investigate and optimize the heat treatment parameters of 17-4
PH stainless steel. This includes standard aging, double aging, and cryogenic treatments followed by aging.
The goal is to enhance both mechanical properties (such as tensile strength, impact toughness, and
hardness) and corrosion resistance through tailored heat treatments.

This study aims to bridge gaps in understanding how unconventional or advanced heat treatments influence
the microstructural evolution and performance of 17-4 PH stainless steel.

2.1.5 Scope of the Literature Review

To contextualize this study, the literature review will cover:

• Metallurgy of 17-4 PH stainless steel: phase transformations, precipitation mechanisms.


• Conventional heat treatments: solution treatment and single-step aging.
• Advanced treatments: double aging (DA), cryogenic aging (CA).
6
• Effect of heat treatments on:
o Microstructure (martensite, delta ferrite, Cu precipitates)
o Mechanical properties (strength, ductility, toughness)
o Corrosion resistance
This comprehensive understanding will support the design and evaluation of experimental heat treatment
schedules and their optimization.

2.2 Metallurgy of 17-4 PH Stainless Steel

2.2.1 Chemical Composition and Alloying Elements

17-4 PH stainless steel, also known as UNS S17400 or SAE Type 630, is a precipitation-hardening martensitic
stainless steel renowned for its high strength and moderate corrosion resistance. Its nominal chemical
composition is as follows:

Table 2.2.1 Nominal Chemical Composition of 17-4 PH

Element Weight %

Chromium (Cr) 15.0 – 17.5

Nickel (Ni) 3.0 – 5.0

Copper (Cu) 3.0 – 5.0

Niobium (Nb) + Tantalum


0.15 – 0.45
(Ta)

Manganese (Mn) ≤1.0

Silicon (Si) ≤1.0

Carbon (C) ≤0.07

Phosphorus (P) ≤0.04

Sulfur (S) ≤0.03

Iron (Fe) Balance

Source: AZoM

7
Chromium (Cr): Chromium is pivotal in imparting corrosion resistance to stainless steels. It forms a passive
chromium oxide (Cr₂O₃) layer on the surface, which protects the underlying metal from oxidative degradation.
This passive film is self-healing, enhancing the material's resistance to various forms of corrosion, including
pitting and crevice corrosion.

Nickel (Ni): Nickel stabilizes the austenitic phase in stainless steels. In 17-4 PH, the nickel content is carefully
controlled to ensure that, upon cooling from the solution annealing temperature, the austenite transforms into
martensite. This transformation is crucial for achieving the desired mechanical properties.

Copper (Cu): Copper is the primary element responsible for precipitation hardening in 17-4 PH stainless
steel. During aging, copper precipitates as fine particles within the martensitic matrix, enhancing strength and
hardness. The precipitation sequence involves the formation of copper-rich clusters, followed by the nucleation
and growth of coherent ε-Cu precipitates.

Niobium (Nb): Niobium acts as a carbide stabilizer, forming niobium carbides (NbC) that prevent the
precipitation of chromium carbides at grain boundaries. This stabilization enhances resistance to intergranular
corrosion and maintains the integrity of the microstructure at elevated temperatures.

Manganese (Mn) and Silicon (Si): Both elements serve as deoxidizers during steelmaking, improving the
cleanliness of the steel and aiding in the removal of oxygen from the melt.

Carbon (C): The carbon content is kept low to enhance weldability and corrosion resistance. Excessive carbon
can lead to the formation of chromium carbides, which deplete chromium from the matrix and reduce corrosion
resistance.

Phosphorus (P) and Sulfur (S): These elements are considered impurities. High levels can lead to the
formation of brittle phases and inclusions, adversely affecting mechanical properties and corrosion resistance.

Fig. 2.2.1: Graph of metals used to produce 17-4 PH SS

8
2.2.2 Phase Transformations

1. Solution Annealing

Solution annealing involves heating the steel to approximately 1040°C (1900°F) and holding it at this
temperature to dissolve existing precipitates and achieve a homogeneous austenitic structure. Subsequent air
cooling transforms the austenite into martensite, setting the stage for precipitation hardening.

Fig. 2.2.2: (a) SEM of As received sample; (b) SEM of Solution Annealed sample

2. Martensitic Transformation

The martensitic transformation in 17-4 PH stainless steel is a diffusionless, shear-type transformation


occurring during cooling. The martensite start (Ms) temperature is approximately 140°C, and the finish (Mf)
temperature is around 30°C. Alloying elements like nickel and copper influence these temperatures .

The resulting martensite has a body-centered tetragonal (BCT) structure with high dislocation density,
contributing to the steel's hardness. However, some austenite may be retained, affecting ductility and
toughness.

Fig. 2.2.3: Evolution of martensitic transformation

9
3. Precipitation Hardening (Aging)

Aging involves reheating the martensitic structure to a lower temperature (typically around 480°C) to
precipitate fine particles that strengthen the steel. The sequence includes:

1. Formation of copper-rich clusters.


2. Nucleation and growth of coherent ε-Cu precipitates.
3. Overaging, where precipitates coarsen, leading to decreased strength.
The aging temperature and time significantly influence the size, distribution, and coherency of the precipitates,
thus affecting mechanical properties.

2.2.3 Microstructural Constituents

Martensite: The primary microstructural constituent in 17-4 PH stainless steel post-solution annealing and
cooling is lath martensite. Its morphology and distribution are influenced by cooling rates and prior austenite
grain size.

Retained Austenite: Some austenite may not transform into martensite during cooling, remaining as retained
austenite. Its presence can enhance toughness but may reduce hardness and dimensional stability.

Delta Ferrite (δ-Ferrite): δ-Ferrite can form during solidification, especially in welds or castings. While it
can improve resistance to hot cracking, excessive δ-ferrite may lead to reduced toughness and anisotropy in
mechanical properties.

Carbides: At higher temperatures or prolonged exposure, carbides such as chromium carbides can precipitate,
especially at grain boundaries, leading to sensitization and reduced corrosion resistance.

Fig. 2.2.4.: SEM micrographs showing the microstructure of 17-4 PH stainless steel, highlighting martensite,
retained austenite, δ-ferrite.

10
2.3 Conventional Heat Treatment of 17-4 PH Stainless Steel

2.3.1 Solution Annealing

Solution annealing is a critical initial heat treatment step for 17-4 PH stainless steel, aiming to homogenize the
microstructure and prepare the alloy for subsequent aging treatments.

Purpose of Solution Annealing:

• Dissolution of Precipitates: Heating the alloy to approximately 1040°C (1900°F) dissolves existing
precipitates, such as copper-rich phases, ensuring a uniform austenitic structure .BOYI
• Stress Relief: This process alleviates residual stresses induced during prior mechanical processing,
enhancing dimensional stability.
• Homogenization: Uniform distribution of alloying elements like Cr, Ni, Cu, and Nb is achieved, which
is essential for consistent mechanical properties.
Temperature Control:

Precise temperature control is paramount.

• Underheating Risks: Temperatures below the optimal range may not fully dissolve precipitates,
leading to heterogeneous microstructures.
• Overheating Risks: Exceeding recommended temperatures can cause excessive grain growth,
adversely affecting mechanical properties.
Grain Size and Mechanical Properties:

Grain size significantly influences mechanical behavior.

• Fine Grains: Enhance strength and toughness.


• Coarse Grains: May reduce ductility and impact resistance.
Standard Industrial Practices:

Industrially, solution annealing is performed at 1040°C for 30 minutes, followed by air cooling. This treatment
is designated as "Condition A”.

11
Fig. 2.3.1: A schematic illustrating the solution annealing
process, highlighting temperature profiles and cooling
rates.

2.3.2 Standard Aging

Standard aging, or precipitation hardening, follows solution annealing to enhance strength and hardness
through controlled precipitation of secondary phases.

Precipitation Hardening Mechanism:

Aging induces the formation of fine, coherent copper-rich precipitates (ε-Cu phase) within the martensitic
matrix. These precipitates hinder dislocation movement, thereby increasing strength and hardness.

Effects of Aging Temperature:

• Low Temperatures (450–480°C): Promote fine, coherent precipitates, resulting in peak hardness and
strength.
• High Temperatures (550–620°C): Lead to coarser, less coherent precipitates, reducing strength but
improving ductility and toughness.
Effects of Aging Time:

• Short Duration: May result in underaging, where precipitates are not fully developed, leading to
suboptimal properties.
• Extended Duration: Can cause overaging, where precipitates coarsen, diminishing strength and
hardness.
Peak Aging:

An optimal combination of aging temperature and time yields peak aging, balancing strength and toughness.

12
"H" Conditions and Corresponding Properties:

Table 2.3.1: Different aging treatments are standardized as "H" conditions

Aging Temp Tensile Strength Yield Strength Elongation Hardness


Condition
(°C) (MPa) (MPa) (%) (HB)

H900 482 1310 1172 10 388–444

H925 496 1172 1069 10 375–429

H1025 552 1069 1000 12 331–401

H1075 579 1000 862 13 311–375

H1100 593 965 793 14 302–363

H1150 621 931 724 16 277–352

Fig. 2.3.1: Graphs showing variation in (a) hardness, (b) ultimate tensile strength

13
2.4 Advanced Heat Treatment Techniques for 17‑4 PH Stainless Steel

Double aging is a two-step age-hardening process in which 17‑4 PH steel is aged at a lower temperature and
then at a higher temperature to tailor precipitate formation and relieve stress. Its objectives include refining
the size and distribution of ε-Cu and Ni precipitates, reducing internal residual stresses, and improving
toughness at the expense of some strength. Early work by Rakhshtadt et al. showed that double aging in
martensitic stainless steels substantially raises the elastic (yield) limit while reducing elastic imperfections. In
practice, double-aging protocols (e.g. the NACE-approved H1150M/DH1150: 760°C→620°C or
620°C→620°C) are used to soften 17‑4PH slightly and enhance ductility without compromising its corrosion
resistance. Recent studies (e.g. Hsieh et al. 2024) have explored various double aging schedules; for example,
a two-stage aging totaling 12 h produced a “soft-tough” condition with UTS ≈900 MPa and elongation ≈26%.

Mechanistically, the first (lower-temperature) aging step nucleates fine Cu-rich precipitates in the martensite
or even austenite (if present), while the second (higher-temperature) step promotes controlled growth or
transformation of those precipitates. For instance, aging in austenite has been shown to form fine Cu clusters
followed by Ostwald ripening of Cu particles. In the martensitic structure, double aging encourages a very
fine, uniform precipitate distribution. SEM images illustrate these effects:

Fig. 2.4.1: SEM micrographs of 17‑4PH steel in different conditions. Left = as-forged (F), center = furnace-
cooled (FC) after solution treatment, right = single-aged at 480 °C (1 h)

14
Fig. 2.4.2: (a) Tensile stress–strain curves and (b) yield
strength (YS), ultimate tensile strength (UTS), and elongation
for as-quenched (F), furnace-cooled (FC), and 480 °C-aged
17‑4PH steel

• Tensile/Yield Strength: Usually slightly lower than the peak single-age condition. Double aging tends
to coarsen some precipitates and lower hardness compared to the maximum-aged state.
• Ductility/Elongation: Greatly increased. The multi-step aging softens the steel; Hsieh et al. reported
26% elongation after double-aging vs. only ~2% in peak-aged steel.
• Toughness: Significantly improved. The increased elongation and finer microstructure translate to
much higher impact energy, making the steel “soft-tough.”
• Corrosion Resistance: Largely maintained. Properly executed double aging in 17‑4PH generally
retains excellent corrosion behavior. (Note: very high aging temperatures can slightly deplete Cr at
grain boundaries, which is why standards restrict 17‑4PH to specific two-step cycles)
The user’s specific schedules – DA1B (480 °C×2 h + 550 °C×4 h) and DA2B (520 °C×2 h + 570 °C×4 h) –
follow the same principles. DA1B’s lower first-stage temperature is expected to produce finer ε-Cu precipitates
and slightly higher initial strength, whereas DA2B’s higher aging temperatures likely coarsen precipitates
15
more (reducing hardness). Both double-aging treatments should lower ultimate strength relative to single-stage
aging but dramatically increase elongation and toughness. In summary, double aging in 17‑4PH trades off
some strength for much higher ductility and toughness, consistent with literature reports and standard practices.

Cryogenic Treatment. Cryogenic treatment (CT) involves cooling 17‑4PH steel to very low temperatures
(typically using liquid nitrogen at –196 °C) and holding for several hours. Traditionally applied to tool, die,
and bearing steels, CT is used to improve hardness, wear resistance, and dimensional stability. During CT,
components are gradually cooled to sub-zero temperatures (often down to –100 to –196 °C) and soaked to
transform metastable phases. These treatments date back to the 1930s in aerospace and tool industries for
“stabilizing” steels. In modern practice, CT may precede or follow standard tempering/aging to exploit these
low-temperature effects.

Microstructurally, CT of martensitic 17‑4PH induces several changes: it transforms retained austenite into
martensite and produces a more refined, lath-like martensitic matrix. The sudden thermal contraction generates
a high density of dislocations, twins, and nanometer-scale defects. Importantly, Jurči and Dlouhý report that
CT markedly reduces retained austenite and refines martensite while dramatically increasing the number of
nanoscale precipitates during subsequent tempering. By analogy, 17‑4PH steel will likely see similar effects:
reduced RA, increased dislocation density, and effectively more nucleation sites for Cu precipitates during
later aging. In summary, cryogenic cooling of 17‑4PH tends to (i) eliminate most RA, (ii) refine martensite
and introduce defects, (iii) alter precipitate kinetics, and (iv) generate extra nanoscale precipitates. These
microstructural changes can improve hardness and wear performance, as commonly observed in steel CT
studies.

Soaking time in CT can be significant. Longer holds (e.g. 7–10 h vs. 4 h) allow more complete martensite
formation and defect accumulation. Some analyses suggest that solute clusters (e.g. carbon) at dislocations
increase with longer CT duration and lower temperature, acting as additional nucleation sites during
subsequent tempering. However, diffusion is essentially frozen at cryogenic temperatures, so any “soak”
effects are indirect. In practice, many reports note that most transformation occurs within the first few hours
of CT.

Reported effects of CT on 17‑4PH properties are mixed. In general, CT often yields a slight hardness and
strength increase (due to extra martensite and defect hardening), but ductility may drop if the steel becomes
more brittle. For example, some studies on similar alloys show minor increases in yield strength after CT at –
196 °C. Toughness trends vary: removing RA can reduce impact resistance, but the refined microstructure can
partly compensate. Wear resistance almost always improves with CT. Corrosion resistance is typically
unchanged or slightly reduced (extra defects can locally accelerate corrosion). Overall, literature on cryogenic
17‑4PH shows both positive and negligible effects. Jurči’s review notes that the mechanisms (e.g. carbon
16
segregation at –196 °C) remain debated, so claimed benefits sometimes lack clear evidence.

Cryogenic Treatment Followed by Aging. Treating 17‑4PH with cryogenic processing prior to aging aims to
combine the benefits of both methods. The rationale is that CT “primes” the steel by eliminating retained
austenite and creating many nucleation sites, so that a subsequent aging step produces a very dense, fine
precipitate population. In effect, CT+aging can dramatically accelerate precipitation hardening kinetics. For
example, Jurči et al. report that steels subjected to CT show a “significantly enhanced number and population
density of nano-sized precipitates” upon tempering. In 17‑4PH, this would translate to more and finer ε-Cu
(and possibly Ni) precipitates after the aging cycle. Furthermore, all martensite formed during CT is available
to be strengthened by aging.

Limited specific studies exist for 17‑4PH CT+aging, but results from analogous steels suggest: strength and
hardness increase beyond aging alone, while the tensile/ductility balance is preserved. For instance, tool-steel
data show that CT followed by tempering yields much higher hardness due to the extra precipitates. By
analogy, 17‑4PH treated with, say, –196 °C soak then aged at 550–570 °C should exhibit higher yield strength
than a purely aged sample, with only modest loss of elongation. Toughness may also improve if the precipitates
are very fine. However, excessive CT (e.g. very long soaks) could slightly reduce impact toughness due to
dislocation embrittlement. Corrosion behavior under CT+aging is not well documented; one might expect little
change or a small decrease if defects increase.

Cryo/aging parameters directly affect the outcome. For the user’s conditions (CT soak for 4, 7, or 10 h
followed by aging), we anticipate that longer CT yields slightly more martensite and defects, thus enabling
more nucleation. Jurči et al. noted that cluster formation at dislocations increases with longer CT time.
Therefore, comparing 4→7→10 h soaks could reveal a trend toward higher strength/hardness and finer
precipitates as soak time increases, until saturation is reached. Any observed strength or hardness gains with
longer soak would confirm that CT-generated nucleation continues accumulating up to that duration.
Conversely, if properties plateau between 7 h and 10 h soaks, this would indicate a limit to the CT-induced
effects. In summary, CT + aging in 17‑4PH is expected to produce finer precipitate distributions (from
enhanced nucleation) and therefore higher strength than aging alone. Varying the cryogenic soak time will
illuminate how quickly these microstructural changes develop, potentially revealing optimal soak durations
for maximum benefit.

17
2.5 Influence of Heat Treatments on 17-4 PH Stainless Steel

2.5.1 Influence on Microstructure

Solution annealing of 17‑4 PH (around 1020–1050 °C) dissolves prior precipitates and yields a martensitic
(body‐centered tetragonal) matrix upon quenching. This results in a fine, recrystallized grain structure and
very high dislocation density. Only a small amount of δ‐ferrite typically remains after solution treatment.
Optical and SEM imaging show the characteristic lath‐martensite structure; TEM reveals that immediately
after quench no Cu‐rich precipitates are present in martensite (only coarse Cu particles appear in any δ‐ferrite).
Subsequent standard aging (in the 480–620 °C range) nucleates very fine, coherent Cu-rich precipitates (~2–
5 nm) in the martensit. With increasing aging time or temperature, these particles coarsen (from uniform
nanoclusters to larger agglomerates. Two‐stage or “double” aging (e.g. a high‐temperature temper followed
by lower‐temperature aging) further coarsens Cu‐precipitates and produces reverted austenite (Ni/Cu-
enriched fcc islands) in the matrix. Cryogenic treatment (deep cooling) of quenched 17‑4 PH converts any
retained austenite to martensite and introduces new lattice defects, effectively refining the martensitic laths
and boosting dislocation density.

• Grain Size: Solution treatment at 1024–1052 °C “refine[s] the grain structure”. Typical prior‐austenite
grains are on the order of 10–20 µm and change little during aging. Thus aging and cryogenic steps
have negligible effect on grain size.
• Precipitate Morphology: Aging at ~480–520 °C produces uniformly dispersed, nanoscale Cu‐rich
precipitates. Higher aging (580–620 °C or long times) leads to precipitate coarsening and
agglomeration. Two‐stage aging (overaging) yields much coarser Cu‐rich particles and also forms soft
reverted-austenite around them.
• Dislocations and Matrix Phases: The as‐quenched martensite has very high dislocation density.
Aging reduces dislocation density (slight recovery) but precipitation hardening compensates by
impeding dislocations. TEM analysis shows dislocations drop off as aging temperature increases. If
any retained austenite were present after quench, deep cryogenic cooling drives its martensitic
transformation (adding dislocations). No new phase forms with aging except reverted austenite at
≥600 °C; δ‐ferrite persists unaltered (it is minor).
• Effects of Cryogenics: Deep-freezing (liquid N₂) of quenched 17‑4PH further reduces retained
austenite and generates refined martensite with more dislocations. This can slightly increase hardness
and alter the subsequent aging response, since more sites for Cu nucleation may form.
Microscopy (OM, SEM, TEM) is routinely used to characterize these changes. Optical microscopy or EBSD
confirms grain/lath morphology (e.g. ~8–9 µm martensitic packets), SEM reveals fracture modes or δ‐ferrite
distribution, and TEM shows nanoscale Cu‐precipitates and dislocation structures. Overall, fine martensite
18
laths with nanoscale Cu‐particles (as in H900/H1025) correspond to maximal strength/hardness, whereas
coarser precipitates and more reverted austenite (as in H1150 or double-aged) yield higher ductility. Uniform
Cr distribution maintains passivity, while any coarse intermetallics or segregation (from extreme aging) can
create local anodic sites that worsen corrosion.

2.5.2 Influence on Mechanical Properties

The mechanical response of 17‑4PH is dominated by its strengthening mechanisms. The ultrafine precipitates
produced during aging impede dislocation motion (precipitation hardening) and the lath‐martensite matrix
itself has high strength due to its high dislocation density. As a result, peak‐aged conditions (H900,
~482 °C/1 h) give the highest strength and hardness. For example, H900 17‑4PH typically has UTS ≈1250–
1400 MPa and yield ≈1100–1300 MPa, with Rockwell C hardness ~43–45. However, this comes at the
expense of ductility and toughness (elongation ~6–8% and Charpy-impact ~15 ft-lb). In contrast, over-aged
conditions (e.g. H1150, ~620 °C/4 h) produce much lower strength (UTS ≈1000 MPa, YS ≈850 MPa) but
greatly improved ductility (elongation ~18–20%) and toughness (impact ~50 ft-lb). A recent gun-barrel study
showed that a two-stage soft-aging (760 °C + 620 °C) yielded UTS ≈900 MPa with 26% elongation (vs.
~1364 MPa/2% in peak H900).

Fig. 2.5.1: Schematic of precipitation strengthening vs particle size.

• Strength–Ductility Trade‐off: Thus there is a trade‐off: H900 (peak strength) ≫ H1150 (toughness).
Rolled Alloys data show H900 condition with ~200 ksi (1379 MPa) UTS, 185 ksi (1276 MPa) YS,
14% elongation, vs. H1150 with ~145 ksi (1000 MPa) UTS, 125 ksi (862 MPa) YS, 19% elongation.
Charpy impact follows the same trend: ~15 ft-lb (H900) vs. ~50 ft-lb (H1150). Double-aging (over-
aging) yields even lower strength and much higher ductility. Cryogenic treatment (after quench) can
further increase yield and hardness slightly by converting residual austenite into dislocation-rich
martensite, though quantitative data on 17‑4PH are scarce.

19
• Strengthening Mechanisms: Key mechanisms include: (1) Martensitic transformation: Quenching
from austenite creates a hard, lath martensite with very high dislocation density. (2) Precipitation
hardening: Nanoscale Cu-rich particles (and niobium carbo-nitrides) precipitate during aging,
blocking dislocations. These precipitates are coherent and most effective at peak aging; as they coarsen
(overaging), they become less effective (Orowan bypass) and strength drops. (3) Dislocation
strengthening: The high dislocation density after quench directly raises strength. Tempering/aging
reduces dislocations, but the simultaneous precipitation compensates. (4) Reverted austenite: Formed
by two-stage aging, this soft phase does not contribute to strength but greatly boosts ductility. (5) Solid‐
solution/other phases: Minor elements (Ni, Nb, Si) also contribute to solid‐solution strengthening and
form stable carbides; however, in 17‑4PH their effect is secondary to Cu-precipitates and martensite.
In summary, the H900 condition maximizes strength via precipitation and lath-martensite hardening, but at
low elongation. H1150 (and double-aging) yields lower strength but much higher toughness, reflecting coarser
precipitates and some reverted-austenite softening. The diagram above illustrates why extremely fine particles
give the best hardening: once precipitates grow beyond a critical size, strength falls.

2.5.3 Influence on Corrosion Resistance

17‑4PH has an inherently good passive film (Cr₂O₃) because of its ~15–17% Cr. In general its corrosion
performance is comparable to Type 304 stainless steel and much better than ferritic 400-series alloys. Heat
treatment affects corrosion indirectly by altering microstructure. In properly aged conditions the Cr is
uniformly distributed, so the passive film remains intact. However, coarse precipitates or intermetallics can
locally deplete alloying elements and create galvanic sites. For example, aggressive aging (high T or long
time) causes Cu-rich particles to coarsen and Ni-rich zones (reverted austenite) to form; these microstructural
heterogeneities promote localized anodic dissolution.

Fig. 2.5.2: Effect of aging on dislocations, precipitates, and corrosion in 17‑4PH

20
• Chromium and Passive Layer: The Cr₂O₃ passive film is stable under all aging conditions. Heat
treatments do not deplete Cr from the matrix, so general corrosion resistance remains high. Slight
differences do arise: mid-range aging (around 550–580 °C) tends to give the best passivity, while
under- or over-aging can worsen it. Zhou et al. showed the corrosion current density drops at 580 °C
aging and then rises at higher aging. This is because moderate aging produces a uniform distribution
of fine Cu-precipitates and low dislocation heterogeneity, minimizing micro-galvanic cells. At higher
aging, precipitate coarsening and elemental segregation increase local corrosion (the study found
580 °C/4 h gave optimal pitting resistance).
• Austenite, Ferrite, and Intermetallics: Retained or reverted austenite (fcc, Ni-rich) has somewhat
different electrochemistry than martensite, but tests on stainless steels indicate that austenite can be
relatively passive (even acting as a barrier to pit growth). In 17‑4PH, any reverted austenite from over-
aging is a tough phase that may also modestly reduce corrosion rates. Delta-ferrite (bcc, Cr-enriched)
if present is similarly passive. The main corrosion concern is precipitates: extremely coarse particles
or carbides (if present) at grain boundaries could initiate intergranular attack, but 17‑4PH has very low
carbon so M₁₃C₆ formation is minimal.
• Environmental Examples: In neutral chloride (salt-spray) environments, 17‑4PH performs like 304
SS. However, peak‐aged (hard) conditions are more prone to chloride stress corrosion cracking (SCC).
ATI data note that “17-4PH is more susceptible to SCC at peak strength,” so chloride-exposed
components are typically aged at the highest temperature that meets strength needs (H1150). Indeed,
NACE guidelines require H1150-D/M for use in sour (H₂S-containing) service. In acid or chloride
media, H900 parts may crack under stress, whereas H1150 parts (with lower hardness) show much
better fracture resistance. Pitting and crevice corrosion susceptibility follow a similar trend: studies
show the highest pitting potentials occur near mid-aging (~580 °C), implying over-aged material is
slightly more pit-prone.
2.5.4 Influence on Fracture Behavior

The fracture mode of 17‑4PH depends strongly on microstructure. In highly aged or martensitic conditions
(H900), fractures tend to be brittle or quasi-cleavage in nature, with little plastic deformation. SEM of broken
specimens typically shows flat facets along martensite laths or packets. In contrast, in softer conditions (H1150
or double-aged with reverted austenite), failure is ductile: surfaces are covered with microvoid‐coalescence
dimples. For example, Zhou et al. observed that their soft-aged 17‑4PH had “a more significant proportion
and size of dimples” on the fracture surface, consistent with its ~26% elongation.

21
• Microstructural Origins: Grain boundaries in 17‑4PH are usually clean (no continuous carbide film),
so intergranular fracture is not common unless severe embrittlement occurs. Instead, cracks nucleate
at inclusions (e.g. sulfides, oxides) or coarse precipitates, then propagate through the martensite.
Nanoscale Cu precipitates do not usually nucleate cracks; rather, they contribute to strength. Reverted
austenite pockets (soft, ductile) can arrest or deflect cracks, increasing toughness. In general, the
transition from brittle (cleavage-like) to ductile (dimpled) fracture correlates with reduced strength:
H900 fractures show fine micro-cleavage, whereas H1150 fractures show deep dimples..
• Correlation with Mechanical Results: The fractography aligns with the tensile/impact data. H900
(UTS ~1300 MPa, low elongation, Charpy ~15 ft-lb) failed suddenly with nearly flat facets. H1150
(UTS ~1000 MPa, high elongation, Charpy ~50 ft-lb) failed by extensive plastic tearing. Double-aged
or softened conditions (760+620 °C) that yielded ~900 MPa UTS and ~26% elongation exhibited fully
ductile fracture with large dimples. Thus, fine-grained tempered martensite favors brittle fracture,
whereas coarse or partly austenitic microstructures yield ductile fracture.

2.6 Characterization Techniques for Heat-Treated 17‑4 PH Stainless Steel

2.6.1 Tensile Testing

Tension testing applies a uniaxial load to a standardized specimen (e.g. round rod) while measuring force and
elongation. The resulting engineering stress–strain curve is plotted (stress = load/initial area; strain =
Δlength/original gauge length). From the curve one finds:

• Ultimate Tensile Strength (UTS) – the maximum stress sustained (peak of curve).
• Yield Strength (YS) – defined by convention (e.g. 0.2% offset) at which plastic deformation begins.
In practice a line parallel to the elastic (Hooke’s) portion is offset by 0.002 strain and its intersection
with the curve gives the 0.2% yield.
• Percent Elongation (%EL) – the increase in gauge length after fracture, divided by original gauge
length (times 100%). ASTM E8/E8M governs tension tests of metallic materials. For example, ASTM
E8 specifies gauge length = 4× diameter (round specimens).
Key factors: the gauge length must be measured precisely (it directly affects % elongation). The strain rate
(crosshead speed or extensometer rate) is controlled (often ≈1–10⁻³ s⁻¹) to ensure consistent results. Higher
strain rates tend to increase measured YS and UTS while reducing elongation. Standard practice (ASTM E8)
also specifies using an extensometer to measure strain within the gauge length and aligning the specimen to
avoid bending stress.

22
Typical output is a stress–strain diagram. The UTS is read at the curve peak. The 0.2% offset yield strength
is found by moving a line 0.2% strain right of the elastic line (Instron definition). Elongation is calculated
from final and original gauge lengths. ASTM E8 (or ISO 6892) provides the test conditions and calculations.

2.6.2 Impact Testing (Charpy)

The Charpy V-notch test (ASTM E23) assesses notch toughness (ability to absorb energy in a rapid fracture).
A standard 10×10×55 mm bar with a 45° V-notch (2 mm deep) is placed on supports and struck by a pendulum
hammer. The absorbed energy (difference in pendulum potential energy before vs. after break) is read from
the tester (in joules). This energy is a measure of impact toughness – higher values indicate a tougher, more
ductile response.

Key points: the notch concentrates stress, forcing crack initiation at a known location. Charpy specimens are
defined in ASTM E23 (with V- and U-notch options). In the Charpy method the specimen is supported
horizontally and hit on the opposite side of the notch. The impact energy is reported (often normalized by
cross-sectional area to yield an impact strength). Results can be correlated to fracture toughness or used to
map ductile–brittle transition behavior (especially in martensitic steels).

Typical procedure: the specimen is loaded at a controlled temperature (room or subzero for transition studies).
The pendulum is released; the post-fracture height indicates energy absorbed. By ASTM E23, test machines
must be calibrated and results reported with specimen orientation and temperature. A Charpy result is usually
given as energy (J) and sometimes as normalized toughness (J/cm²).

2.6.3 Hardness Testing

Hardness measures a material’s resistance to indentation. Two common methods were used: Rockwell C
(macro-hardness) and Vickers microhardness. In Rockwell C (ASTM E18), a 120° diamond cone indenter
(brale tip) is loaded with 150 kgf total. The hardness number (HRC) is determined from the depth of the
permanent indent: a larger indentation (deeper) gives a lower hardness. Rockwell testing yields a direct readout
(no optical measurement) and is fast for bulk hardness.

Vickers microhardness (ASTM E384/E92) uses a diamond pyramid indenter (136° included angle) pressed
into the polished surface under a small load (e.g. 100 gf – 1 kgf). After the dwell time, the two diagonals of
the square impression are measured under a microscope, and hardness is calculated.

In both tests, well-prepared flat surfaces and clean conditions are essential. The results on 17‑4 PH were
compared across heat treatments. Typically, a higher hardness indicates higher strength (due to work hardening
or precipitation). Indeed, empirical correlations show UTS ≈ (3–3.5)·HRC in many steels. Thus hardness
testing is a quick proxy for strength.
23
2.6.4 Microstructural Analysis

Fig. 2.6.1: Optical micrograph of a martensitic stainless


steel (AISI 4140) showing a lath martensite
microstructure (etched).

Optical microscopy reveals the grain and lath structure. Samples are metallographically polished and etched
(e.g. with a suitable etchant for martensitic SS) per ASTM E3. The etched grain or prior-austenite boundaries
and martensite laths are observed under ×100–×500 magnification. Grain (or lath packet) size can be measured
(ASTM E112) by line-intercept or comparison charts. For 17‑4 PH, the microstructure is primarily tempered
martensite. As one source notes, 17‑4 PH contains mostly martensite and (depending on heat treatment) some
retained austenite or δ-ferrite, along with fine precipitates (Cr/Nb carbides and Cu-rich particles). Thus optical
micrographs show elongated martensite laths within prior-austenite grains; heat treatment (e.g. aging) can
refine or coarsen these features.

• SEM analysis: Scanning electron microscopy was used for higher-resolution imaging. In
backscattered or secondary-electron mode, SEM can image the fracture surface or reveal fine
precipitates. For example, fractured tensile specimens were sputter-coated and examined by SEM to
characterize fracture morphology (see Fracture Analysis below). SEM can also show precipitate
morphology or carbide distributions, although chemical analysis (EDS) is often needed for
identification. (No TEM was available in this project, but if used, transmission electron microscopy
could image nanoscale precipitates and dislocations – valuable for PH steels – albeit beyond this work.)
• ASTM Standards: Optical metallography follows guides like ASTM E3 (sample prep) and E407
(etching), and grain-size per E112. SEM use is guided by standard practice (e.g. using EDS as per
ASTM E1508). These reveal microstructural differences due to aging or cryogenic treatment (e.g.
precipitate density, lath refinement).
X-ray Diffraction (XRD)
24
XRD identifies crystalline phases via Bragg’s law (nλ = 2d sinθ). Monochromatic X-rays diffract off lattice
planes; peaks at angles θ correspond to interplanar spacings d. In practice, 2θ vs intensity patterns were
collected. The positions of diffraction peaks were matched to reference patterns: martensite (body-centered
tetragonal/orthorhombic peaks) and austenite (face-centered cubic γ-Fe) have distinct signatures. For example,
the (211) martensite peak and (220) austenite peak can be identified. Precipitate phases (e.g. Cu or carbides)
may also produce weak peaks if coarse enough; at least, peak shifts can suggest lattice strain from solute.

From XRD data one can estimate: retained austenite fraction by comparing integrated intensities of γ vs α
peaks. As Pulstec notes, the ratio of the martensite(211) to austenite(220) intensities yields phase volume
fractions. Lattice parameters of phases can be calculated using Bragg’s law from the 2θ positions, revealing
e.g. carbon expansion of martensite. Peak broadening analysis (Scherrer equation) gives average crystallite
size or microstrain: broader peaks imply finer grains or high dislocation density. In summary, XRD confirmed
the predominance of martensite with a minor retained-γ phase (especially after standard aging), consistent with
the known 17‑4 PH microstructure.

2.6.5 Fracture Analysis (SEM)

Fig. 2.6.2: SEM image of a ductile fracture surface


(aluminum alloy).

After tensile testing, the fracture surfaces of broken specimens were examined by SEM. This fractography
reveals the failure mode:

• Ductile fracture: Characterized by a dimpled (microvoid) surface. Under SEM, many hemispherical
voids (from particle/decohesion) are visible. Void coalescence leaves rough, fibrous topography. Such
ductile, “cup-and-cone” appearance indicates plastic deformation prior to fracture.

25
• Brittle fracture: Characterized by flat facets and cleavage planes. SEM shows cleavage steps or river
patterns with little plastic flow. Brittle surfaces are relatively smooth and often intergranular or
transgranular, depending on path.
• Mixed-mode: Often both features appear. Regions of dimples may surround localized cleavage
features, reflecting partial plasticity.
Interpretation follows classic fractography: fully ductile (all dimples) implies high toughness; brittle features
(planar facets) imply low toughness/embrittlement. In 17‑4 PH, one might see dimpled facets at room temp,
shifting to more cleavage at lower temp. SEM imaging (500–2000×) distinguished these modes.

2.6.6 Corrosion Testing

Potentiodynamic polarization (ASTM G59) was used to evaluate corrosion behavior. In a three-electrode
cell, the steel sample is the working electrode, with a reference electrode (e.g. Ag/AgCl) and a platinum
counter-electrode. The solution (e.g. saline electrolyte) is deaerated or as required. After stabilizing at open-
circuit potential (OCP ≈ corrosion potential E₀), the potential is swept (typically ±250 mV around OCP) at a
slow scan rate (∼0.5–1 mV/s). The resulting current–potential curve (log current vs. E) shows Tafel regions
(linear in log scale).

From this curve, Tafel extrapolation is used to find corrosion parameters. The linear (butterfly-shaped) anodic
and cathodic branches are extrapolated back to their intersection: the intersection current density is the
corrosion current I₍corr₎, and the corresponding potential is E₍corr₎. Physically, E₍corr₎ is the potential where
anodic metal dissolution equals cathodic reduction (no net current), and I₍corr₎ relates to uniform corrosion
rate (by Faraday’s law).

Typical setup details: the scan range often spans from about –0.5 V to +0.5 V versus a reference, relative to
the OCP; standard cell geometry and solution aeration are as per ASTM G59/G5. Data analysis (often software
fits) yields the linear Tafel slopes and intercept. The corrosion rate (mm/year or mpy) can be calculated from
I₍corr₎ assuming metallic density.

Limitations: This method assumes uniform corrosion kinetics and ignores localized effects (pitting). The
extrapolation is only valid if clear Tafel regions exist. Large scan rates or surface films can distort results.
Moreover, polarization testing modifies the surface and electrolyte conditions (it is not a true steady-state
measurement). Thus I₍corr₎ values are comparative and must be interpreted cautiously. Nevertheless,
potentiodynamic polarization is widely used to assess relative susceptibility and to determine parameters
(E₍corr₎, I₍corr₎, passive current) for materials under given conditions.

26
ASTM Standards: ASTM G59 (or G102) covers the method for potentiodynamic polarization and calculation
of corrosion rates. ASTM G5 specifies reference electrodes. These guided the test procedure here.

Significance: The extracted E₍corr₎ indicates thermodynamic tendency (more negative → more active). The
I₍corr₎ (and derived corrosion rate) indicates kinetic aggressiveness. Comparing treatments, a lower I₍corr₎ for
cryo- or double-aged samples would suggest improved corrosion resistance. Tafel analysis is a standard way
to quantify these effects for 17‑4 PH variants.

2.7 Research Gaps and Motivation

• Limited study of double aging and cryogenic time: Most prior work on 17-4PH focuses on single-
step aging (e.g. H900 or H1150). Few systematic studies examine double aging (two-stage aging)
effects, or the influence of different cryogenic soak durations on both strength/toughness and corrosion.
For example, Hung et al. report that a 12 h double-aging (“soft‐tough”) treatment gives ~900 MPa
UTS and 26% elongation, but broader trends are unclear. Similarly, while cryogenic soaking can refine
martensitic steels, the effect of varying times (e.g. 4, 7, 10 h at –196°C) before aging is largely
unreported.
• Novelty of combined treatments: The present study will combine cryogenic pre-treatment with
conventional aging and double aging, systematically comparing multiple cryogenic soak times. No
prior work has applied a comprehensive set of hold times and post-aging steps. Our approach will use
advanced microscopy and mechanical/corrosion tests to link processing to microstructure and
performance.
• Industrial relevance: Tailoring 17-4PH properties via heat treatment is vital for demanding
applications. For instance, 17-4PH is used in oilfield tools, nuclear components, aerospace hardware
and marine structures that require high strength and corrosion resistance. Optimizing toughness-
versus-strength (e.g. for flight-critical parts) and improving corrosion resistance (e.g. for seawater
service) can extend service life in these sectors.
• Research questions:
o Does double aging yield a better toughness–strength balance? (Beyond the single double-age
example, can a two-step aging produce a softer, tougher condition?)
o How does cryogenic soak time affect microstructure and precipitate formation? (Longer
subzero holds may convert more retained austenite to martensite and precipitate fine carbides.)
o Can cryogenic pre-treatment improve corrosion resistance? (By refining the microstructure or
altering precipitates, cryo may affect passive film stability – an open question for 17-4PH.)

27
2.8 Conclusion of Literature Review

Precipitation hardening in 17-4PH produces a trade-off between strength and toughness. For example,
commercial data (Table above) show that the H900 condition (482°C aging) yields very high strength
(~190 ksi UTS) with modest elongation, whereas a double-aged condition (H1150M) drops UTS to ~125 ksi
while boosting ductility to ~22%. In general, aging induces fine Cu‐rich precipitates that strengthen the
martensite matrix, but excessive aging causes martensite lath coarsening and reduced toughness. Cryogenic
treatment refines this microstructure further: subzero soaking transforms retained austenite into fresh
martensite and promotes nanoscale carbide formation, reducing martensite lath width and grain size. These
microstructural changes (higher dislocation density, finer grains/precipitates) are known to increase hardness
and strength in other martensitic alloys but their net effect on 17-4PH corrosion has not been fully explored.

Controlling the sequence of solution treatment, cryogenic hold, and aging is thus critical for tailoring 17-4PH
properties. The literature shows that higher aging temperatures improve corrosion resistance (e.g. 580°C aging
gave much lower corrosion current than 480°C), but the role of cryogenic pre-treatment in passivation remains
unknown. In summary, key findings are: aging precipitates raise strength (via fine Cu phases) but risk
embrittlement if overdone; cryogenic processing can substantially refine martensitic microstructure; and few
studies have examined these effects in combination. We will address these gaps by solution-treating 17-4PH
and then applying single-aging, double-aging, and cryogenic-plus-aging cycles with varied soak times.
Mechanical testing (hardness, tensile/toughness) and corrosion assays will be used alongside microstructural
characterization (SEM/TEM, XRD) to answer the questions above and optimize the heat treatment for high-
performance applications.

28
2.9 Experimental Flowchart

29
CHAPTER 3
EXPERIMENTAL METHODOLOGY

3.1 Material Procurement

For the present study, the material selected was a commercially available cylindrical bar of 17-4 PH stainless
steel. The bar had a diameter of 12 mm and a total length of 6 meters. It was procured in the solution-annealed
state, which provides a uniform and stable microstructure suitable for subsequent heat treatment and cryogenic
processing. This condition is preferred for experimental investigations as it serves as a standard baseline for
evaluating the effects of various thermal treatments on material properties.

The bar was inspected upon delivery to ensure dimensional consistency and absence of surface defects.
Sections were cut from the bar as per the size requirements of various mechanical and microstructural tests,
including tensile testing, hardness measurements and metallographic analysis. The uniformity in material
composition and condition was critical to maintain consistency in results across all experimental procedures.

Fig. 3.1: Samples before heat treatment

3.2 Chemical Analysis of Sample

To confirm the chemical composition and ensure that the received material met the standard specifications of
17-4 PH stainless steel, a detailed chemical analysis was carried out. The objective was to verify the presence
and proportion of key alloying elements such as chromium (Cr), nickel (Ni), copper (Cu), niobium (Nb), and
others, which are critical in determining the alloy’s mechanical and corrosion-resistant properties. A small
section of the cylindrical bar was taken for analysis, and the surface was cleaned thoroughly to remove any
contaminants that might interfere with accurate readings.

30
The analysis was performed using an Optical Emission Spectrometer (OES), a reliable and widely used
technique for elemental quantification in metallic samples. The sample was struck with a high-voltage spark,
causing the atoms in the material to emit characteristic light wavelengths corresponding to each element. The
spectrometer detected and quantified these emissions, producing a detailed report of the elemental
composition. The results were then compared with standard ranges for 17-4 PH stainless steel to confirm
material authenticity and suitability for the intended cryogenic and heat treatment experiments.

Table 3.1: Chemical Analysis

Sr. no. PARAMETER COMPOSITION (%)


1. Carbon 0.065
2. Silicon 0.325
3. Manganese 0.714
4. Phosphorous 0.035
5. Sulphur 0.0083
6. Chromium 15.27
7. Nickel 4.514
8. Copper 3.147

3.3 Heat Treatment Cycles

In order to investigate the influence of thermal and cryogenic treatments on the mechanical and microstructural
properties of 17-4 PH stainless steel, three distinct heat treatment routes were designed and executed: standard
aging, double aging, and cryogenic treatment. All samples across these three batches underwent an initial
solution annealing step, which served as a common baseline to ensure microstructural uniformity prior to
applying further treatment variations. Solution annealing was carried out at 1040°C for a duration of 30
minutes, followed by air cooling. This process dissolves the precipitates and homogenizes the matrix, enabling
effective transformation during subsequent treatments.

Batch 1 was subjected to standard aging, a conventional process widely used for hardening precipitation-
strengthened alloys. After solution annealing, the samples were aged at 480°C for 1 hour. This temperature-
time combination allows for the controlled precipitation of copper-rich particles within the martensitic matrix.
These precipitates act as obstacles to dislocation motion, thereby improving the strength and hardness of the
alloy. The purpose of this batch was to serve as a reference point against which the more advanced treatments
could be compared in terms of mechanical response and microstructural refinement.

31
Batch 2 involved a double aging treatment, which is known to enhance precipitation kinetics and thermal
stability. This batch was further divided into two groups. Group 1 was aged at 480°C for 2 hours followed by
a second aging step at 550°C for 4 hours. Group 2 underwent aging at 520°C for 2 hours followed by 570°C
for 4 hours. The double aging approach allows the formation of a finer and more uniform distribution of
precipitates by initially nucleating them at a lower temperature and then growing them under controlled
conditions at a higher temperature. This strategy aims to improve the alloy’s strength, ductility, and wear
resistance simultaneously, while also mitigating the risk of over-aging, which could reduce mechanical
performance.

Batch 3 focused on cryogenic treatment, a deep-cooling process introduced after solution annealing to enhance
hardness, dimensional stability, and fatigue resistance. After annealing, this batch was divided into three
groups based on cryogenic exposure duration. The samples were cooled to −196°C using liquid nitrogen and
held for 3, 6, and 9 hours respectively for Groups 1, 2, and 3. Cryogenic treatment transforms retained austenite
into martensite and also promotes a denser and more refined microstructure. Extended exposure to cryogenic
temperatures is expected to increase the extent of transformation and potentially improve wear and fatigue
properties, especially in applications involving cyclic or impact loading.

Through this structured approach to heat treatment, the experimental design enables a comparative evaluation
of each method's influence on the alloy's performance. By analysing the mechanical test results and
microstructural features of each treated group, the study aims to identify an optimized treatment route that
balances strength, toughness, and wear resistance. This systematic variation of treatment parameters forms the
foundation for understanding the complex interactions between temperature, time, and cryogenic exposure in
precipitation-hardenable stainless steels like 17-4 PH.

Table 3.2: Standard Aging

Process Parameters

1040°C for 30 min, Air


Solution Annealing
Cooling

Aging Treatment 480°C for 1h

32
Fig. 3.2(a): After annealing in furnace Fig. 3.2(b): Air cooling after annealing

Table 3.3: Cryogenic Treatment

Process Parameters

Solution Annealing 1040°C for 30 min, Air Cooling

Cryogenic Treatment -196°C for 3h, 6h, 9h

Aging Treatment 480°C for 1h

33
Fig. 3.3(a): Liquid Nitrogen poured Fig. 3.3(b): Cryogenic treatment
into chamber chamber

Fig. 3.3(c): After cryogenic


treatment

Table 3.4: Double Aging

Process Parameters

Solution Annealing 1040°C for 30 min, Air Cooling

Double Aging 1st Aging: 480°C-520°C for 2h

Treatment 2nd Aging: 550°C-570°C for 4h

34
3.4 Microstructural Evolution

Microstructural analysis is a fundamental step in evaluating the impact of different heat treatment cycles on
the internal structure of 17-4 PH stainless steel. While mechanical testing offers quantitative data on strength,
hardness, impact resistance, and corrosion behaviour, it does not reveal the internal microstructural
mechanisms responsible for these variations. Therefore, observing the microstructure at the microscopic level
becomes essential to correlate the heat treatment process with changes in material performance.

In this study, optical microscopy was used to examine the microstructural evolution of samples after standard
aging, double aging, and cryogenic treatment. Samples were polished and etched using standard
metallographic procedures, and images were captured at magnifications of 100X and 400X. These
magnifications allow for the clear observation of grain morphology, phase distribution, and qualitative
differences in precipitate dispersion or refinement across treatment conditions. Even though optical
microscopy does not reveal nanoscale features, it remains highly effective for identifying martensitic
transformation, grain boundary definition, and microstructural uniformity—especially useful for comparison
between different treatment batches.

Fig. 3.4: Optical Microscope

Through microstructural observation, it becomes possible to assess the extent of phase transformation induced
by each treatment route. For instance, changes in the appearance or density of martensitic laths, differences in
grain coarsening, or the presence of phase segregation can offer visual evidence of how specific treatment
parameters, such as aging temperature or cryogenic exposure duration, influence the internal structure. These
observations help to explain the trends noticed in mechanical test results, such as variations in tensile strength,
impact toughness, hardness, and corrosion behaviour.

Ultimately, the goal of conducting microstructural analysis is to complement mechanical testing and develop
a deeper understanding of how the thermal history affects the alloy's internal characteristics. Observing these
35
samples at multiple magnifications enables a more thorough qualitative assessment of structural evolution
across treatments. The findings contribute significantly to drawing correlations between processing conditions
and material performance, guiding the optimization of heat treatment parameters for enhanced application-
specific behaviour in 17-4 PH stainless steel.

To ensure accurate and high-quality microstructural analysis under the optical microscope, proper sample
preparation is essential. The metallographic preparation began with sectioning of the heat-treated samples into
smaller, manageable pieces suitable for mounting and polishing. The cut samples were carefully ground using
successive grades of silicon carbide (SiC) abrasive papers, starting from coarse (e.g., 220 grit) and progressing
to finer grits (up to 1200 grit), under continuous water flow to avoid overheating and surface deformation.

After initial grinding, lapping was performed using colloidal silica, which is a critical step for achieving a
scratch-free, mirror-like finish on the sample surface. Colloidal silica, a fine suspension of amorphous silica
particles in water, acts as both a lubricant and a polishing agent. The samples were placed on a soft polishing
cloth saturated with colloidal silica and lapped for an extended duration under low pressure. This final
polishing step is particularly effective for revealing fine structural details by minimizing surface damage and
deformation introduced during earlier grinding stages.

Following polishing, the samples were rinsed thoroughly with distilled water, followed by ethanol, and then
dried using warm air. The next step was chemical etching using Kalling’s Reagent No. 2 to reveal the
microstructure under optical magnification. The etchant was freshly prepared with the following composition:
5 g of copper (II) chloride (CuCl₂) dissolved in 100 mL of concentrated hydrochloric acid (HCl), and then
mixed with 100 mL of ethanol. The etched samples were immersed in the reagent for a period ranging from 8
to 12 seconds, depending on the sample and the level of contrast required. Etching time was carefully
controlled to avoid over-etching or inadequate development of structural features.

After etching, the samples were immediately rinsed in water, then ethanol, and gently dried using a stream of
warm air. The prepared specimens were then examined under an optical microscope at 100X and 400X
magnifications to observe the general grain structure, martensitic regions, and any differences in
microstructural features that resulted from the various heat treatment processes. This preparation procedure
ensured clarity and consistency across all samples, enabling reliable comparative analysis.

36
Table 3.5: Kaling's Reagent 2

COMPONENT QUANTITY
Copper Chloride (CuCl2) 5g
Hydrochloric acid (concentrated) 100ml
Ethanol (concentrated) 100ml

3.5 Rockwell Hardness Testing

Hardness is a critical mechanical property that reflects a material's resistance to localized plastic deformation,
typically measured through indentation techniques. Among various hardness testing methods, the Rockwell
hardness test is widely used due to its simplicity, speed, and ability to provide direct hardness values without
requiring microscopic measurement of the indent. It is particularly suitable for metals and alloys, including
heat-treated stainless steels like 17-4 PH.

The Rockwell hardness test operates on the principle of measuring the depth of penetration of an indenter
under a specified load. In this method, a minor load is first applied to establish a reference point, followed by
a major load. The difference in depth between the reference and the final load is automatically converted into
a hardness number, corresponding to a specific Rockwell scale.

In the present study, the Rockwell C scale (HRC) was selected over other Rockwell scales because it is
specifically designed for testing harder materials, particularly steels that have undergone heat treatment
processes like aging or cryogenic tempering. Softer scales such as Rockwell B (HRB), which use a steel ball
indenter and lower loads, are not suitable for accurately assessing the hardness of hardened materials like 17-
4 PH stainless steel. The C scale, on the other hand, uses a diamond indenter and a higher load, which provides
better resolution and accuracy for the hardness range expected in this alloy after treatment.

A diamond cone indenter (brale type) with a cone angle of 120° and a tip radius of 0.2 mm was used in
conjunction with the C scale. The test procedure began with the application of a minor load of 10 kgf, which
ensured proper seating of the indenter and eliminated surface irregularities. After a short stabilization period,
a major load of 150 kgf was applied, bringing the total load to 160 kgf during the main measurement phase.

The dwell time—the duration for which the major load was maintained—was set to 10 seconds, allowing the
material to respond fully under the load and ensuring consistent penetration. After the dwell period, the major
load was gradually released while retaining the minor load, and the permanent depth of indentation was

37
measured. The Rockwell hardness value was directly read from the dial or digital display of the hardness
testing machine.

To ensure accuracy and repeatability, multiple readings were taken at different locations on each sample
surface. The average of these values was then considered as the final Rockwell hardness for that particular
sample. This methodology allowed for consistent, efficient, and comparative evaluation of the hardness across
all samples subjected to different heat treatment conditions. The obtained results were later correlated with
microstructural observations and other mechanical properties such as tensile strength, impact toughness,
microhardness, and corrosion resistance, to provide a comprehensive understanding of the effect of thermal
treatments on 17-4 PH stainless steel.

Fig. 3.5(b): Indentation after testing

Fig. 3.5(a): Rockwell


Hardness Tester

38
3.6 Microhardness Testing

The microhardness test is a highly precise method used to measure the hardness of very small regions within
a material, especially useful for characterizing the effects of localized treatments such as heat or cryogenic
processes. Unlike macrohardness tests that assess bulk properties, microhardness testing is capable of detecting
subtle variations across different microstructural phases, grain boundaries, or heat-affected zones. This makes
it particularly suitable for evaluating the fine-scale hardness changes induced by the heat treatment cycles
applied to 17-4 PH stainless steel in this study.

In this investigation, microhardness measurements were conducted using the Vickers microhardness testing
method, which employs a diamond-shaped (pyramidal) indenter. The indenter has a square base and an angle
of 136° between opposite faces. This geometry is preferred for microhardness analysis because it creates a
well-defined, small-scale indentation, making it easier to assess hardness in localized microstructural regions
without significantly damaging the sample surface. The test is especially valuable for evaluating hardness
variations within martensitic, austenitic, and precipitate-rich areas that may not be evident through Rockwell
hardness testing.

A test load of 1000gf (1 kgf) was applied to the sample surface during the test. This load was carefully chosen
to be high enough to penetrate the treated surface layers while remaining within the microhardness testing
range. A dwell time of 15 seconds was maintained to ensure uniform loading and full indentation development.
The load was applied gradually to avoid overshooting or introducing any impact effects. After removing the
load, the size of the diamond-shaped impression was measured using an integrated optical microscope.

Measurements were performed at a magnification of 100X, which allowed for precise focusing on the indent
and accurate measurement of the diagonal lengths. These measurements were then used to calculate the
Vickers Hardness Number (VHN) using the standard formula that accounts for the load and the average
diagonal length of the indentation. To ensure reliability, multiple indentations were made on each sample in
different microstructural regions, and an average value was taken to represent the overall microhardness for
that condition.

This testing methodology provided critical insights into how various treatments—such as standard aging,
double aging, and cryogenic treatment—impacted the localized hardness of the material. By analyzing the
microhardness data in conjunction with optical micrographs and mechanical test results, the study aimed to
draw correlations between thermal history, microstructural evolution, and mechanical performance of 17-4 PH
stainless steel.

39
Fig. 3.6: Microhardness reading

3.7 Testing of Charpy Impact Energy

The Charpy impact test was conducted to assess the toughness and impact resistance of the 17-4 PH stainless
steel samples subjected to various heat treatments. This test provides valuable insight into how a material
absorbs energy during fracture, which is especially critical for components expected to undergo sudden or
dynamic loading. Evaluating the impact energy helps in understanding the effect of different thermal
treatments on the material’s ductility and ability to withstand shock loading.

Sample preparation for the Charpy test was carried out following standard procedures. The samples were
machined into standard V-notched specimens with overall dimensions of 55 mm in length and 10 mm in
thickness, conforming to the requirements of the Charpy impact test. In some cases, sub-size specimens of 7.5
mm or 5 mm thickness were also used where necessary. The V-notch was precisely milled to ensure
consistency in crack initiation across all test specimens. The notch had a depth of 2.0 mm, an angle of 45°,
and a root radius of 0.25 mm. To maintain symmetry, the notch was cantered exactly 27.5 mm from each end
of the specimen, ensuring accurate placement for uniform loading conditions.

The test was performed on a standard pendulum-type Charpy impact testing machine. The pendulum was
raised to a fixed height and then released to strike the notched specimen, which was positioned horizontally
with the notch facing away from the striker. The energy absorbed by the sample during fracture was recorded
by the machine in joules. This energy represents the material’s ability to absorb mechanical shock and is an
indicator of its toughness. The broken pieces were also retained for visual inspection of fracture surface
characteristics.

40
All seven heat treatment variants were tested, and for each condition, three samples were evaluated to ensure
repeatability and accuracy. The average of the three readings was taken as the representative value of impact
toughness for that specific treatment. This helped in eliminating anomalies due to any experimental
inconsistencies or material defects.

The results from the Charpy test were later compared across the different treatment conditions to understand
the influence of standard aging, double aging, and cryogenic treatments on impact energy. Treatments that
improved the distribution or refinement of martensitic or precipitation phases generally showed better
toughness values. On the other hand, treatments that led to embrittlement or excessive phase precipitation
showed reduced impact energy absorption.

The Charpy impact data was ultimately used alongside other mechanical test results such as tensile strength
and hardness to gain a more comprehensive understanding of the material behaviour. These findings contribute
toward optimizing the heat treatment cycle for 17-4 PH stainless steel to achieve a desirable balance between
strength and toughness for engineering applications.

Fig. 3.7(a): Impact machined Fig. 3.7(b): Sample after impact


sample testing

41
3.8 Determination of YS, UTS, %Elongation
The tensile test is a standard mechanical characterization method used to determine how a material behaves
when subjected to a uniaxial pulling force. It provides valuable information such as yield strength, ultimate
tensile strength (UTS), and percentage elongation, all of which are essential parameters to evaluate the
mechanical performance of metallic materials after undergoing various thermal treatments. In the current
study, tensile testing was used to assess the strength and ductility variations in 17-4 PH stainless steel
following standard aging, double aging, and cryogenic treatment cycles.

The tensile specimens were prepared from a 12 mm diameter cylindrical bar of 17-4 PH stainless steel, which
had been subjected to various heat treatment conditions. Standard round-type tensile samples were machined
on a lathe as per ASTM E8 specifications. The gauge section, shoulders, and gripping ends were precisely
turned to maintain dimensional accuracy. A smooth surface finish in the gauge area was ensured to prevent
premature failure from surface imperfections. Maintaining consistency in specimen dimensions was critical to
allow for reliable comparative analysis of mechanical properties across different treatment groups.

To enhance gripping efficiency and prevent slippage during the test, knurling was performed on both ends of
each specimen, outside the gauge length. This involved creating a fine patterned texture using a knurling tool
on the lathe machine. Knurling increased surface roughness in the gripping region, thereby improving the
mechanical interlock between the sample and the jaws of the testing machine. This was especially important
due to the high tensile forces involved and helped ensure accurate load application without sample movement.

The tensile testing was conducted on a 60-ton capacity universal testing machine (UTM), capable of handling
large loads with precision. The prepared specimens were mounted securely between the upper and lower
crossheads of the machine. The test was carried out under room temperature conditions. The machine gradually
applied a tensile load to the specimen at a constant strain rate until fracture occurred. Load and elongation data
were recorded continuously during the test using a digital data acquisition system integrated with the UTM.

From the recorded data, yield strength was determined as the stress at which the material begins to deform
plastically, typically identified using the 0.2% offset method. The ultimate tensile strength (UTS) was
identified as the maximum stress sustained before necking began. The percentage elongation was calculated
by measuring the increase in gauge length after fracture, indicating the ductility of the material. For each of
the seven heat treatment conditions, three tensile specimens were tested to ensure reliability and reproducibility
of results.

The average values of yield strength, UTS, and elongation were computed for each treatment group and were
later used to create comparative plots. These plots provided a visual and quantitative comparison of how
42
different heat treatment strategies influenced the mechanical properties of 17-4 PH stainless steel. The tensile
data, when interpreted alongside microstructural and hardness results, contributed significantly to
understanding the correlation between heat treatment parameters and the material’s overall mechanical
behaviour.

Fig. 3.8(b): Sample after tensile test


Fig. 3.8(a): Machined
Sample before test

Fig. 3.8(c): Fracture of sample


after tensile test

43
3.9 X-Ray Diffractometer Testing

X-ray Diffraction (XRD) analysis is a powerful non-destructive technique widely used for identifying the
crystallographic structure, phase composition, and internal stresses within metallic samples. In this project,
XRD testing was carried out to investigate the phase changes occurring in 17-4 PH stainless steel samples
after undergoing different heat treatment cycles such as standard aging, double aging, and cryogenic treatment.
The test aimed to confirm the presence of martensitic, austenitic, or any secondary phases and to understand
how heat treatment influenced the microstructure at the atomic level.

The samples used for XRD testing were cylindrical in shape with a diameter of 12 mm and a length of 10 mm.
These dimensions were selected to ensure that the specimen surface was wide and flat enough to allow accurate
exposure and signal detection during the test. Each sample was properly cleaned and polished before testing
to remove any surface contaminants, oxides, or machining marks, which could otherwise interfere with the
diffraction pattern and affect the clarity and accuracy of the results.

The XRD analysis was carried out using a standard powder diffractometer equipped with a Cu-Kα radiation
source, which is commonly used due to its high penetration and clear peak resolution in metallic materials.
The X-rays were directed at the polished surface of the cylindrical sample at a fixed incident angle, and the
detector scanned across a range of 2θ angles to detect the diffracted beams. The scanning range typically
covered 20° to 100° 2θ, with an appropriate step size and scanning speed to ensure a good balance between
resolution and testing time.

During the analysis, the interaction of the X-rays with the atomic planes in the crystal structure produced
diffraction peaks at specific angles, which were recorded to create a diffraction pattern. The positions and
intensities of these peaks were then used to identify the phases present in the material by comparing them with
standard reference data. In the case of 17-4 PH stainless steel, particular attention was paid to peaks
corresponding to martensite, retained austenite, and any precipitate phases such as copper-rich or chromium
compounds, which are known to influence mechanical properties.

The diffraction data obtained from each sample were analyzed using specialized software to refine the peak
positions and match them with database patterns. This helped in determining both qualitative and, to some
extent, semi-quantitative phase composition. The presence of any shift in peak position or change in peak
width was also studied, as these could indicate lattice strain or crystallite size variations due to different heat
treatments. Although the samples were not subjected to Rietveld refinement or advanced quantification, the
basic peak analysis provided significant insight into the structural evolution.

44
The XRD results played a critical role in supporting the findings from mechanical testing and microstructural
evaluation. By confirming the phase transformations associated with different aging and cryogenic treatments,
the XRD analysis provided a deeper understanding of the changes occurring within the steel matrix. These
results helped to explain variations in hardness, strength, and ductility observed across the various heat
treatment batches and added scientific support to the conclusions drawn from other experiments.

Fig. 3.9: X-ray Diffractometer

3.10 Corrosion Resistance Testing (Tafel Extrapolation)

Corrosion resistance testing was conducted to assess the electrochemical behaviour of 17-4 PH stainless steel
samples subjected to different heat treatment processes. The testing employed the potentiodynamic
polarization method, specifically the Tafel extrapolation technique, which is effective in evaluating corrosion
tendencies in chloride-rich environments like seawater or industrial applications.

Cylindrical samples of 12 mm diameter and 10–15 mm length were prepared for each of the seven treatment
conditions, including the base material, standard aging, two double aging conditions, and three cryogenic
treatments. One circular face of each sample was designated as the working electrode surface. This surface
was ground using silicon carbide (SiC) abrasive papers of increasing fineness (240, 400, 600, 800, and 1200
grit) and polished using 0.3 µm alumina slurry to achieve a mirror-like finish. The prepared samples were then
rinsed with distilled water, cleaned ultrasonically in ethanol or acetone for 5–10 minutes, and dried with warm
air or tissue to remove contaminants.

45
The electrochemical cell used for testing followed a standard three-electrode setup, with the mounted 17-4 PH
sample serving as the working electrode (WE), a saturated calomel electrode (SCE) or Ag/AgCl electrode as
the reference electrode, and a platinum or graphite rod as the counter electrode. The electrolyte used was a 3.5
wt% NaCl solution, which simulates a typical marine or corrosive environment. The electrolyte volume ranged
from 250–500 mL, depending on the cell size.

Prior to the polarization scan, samples were immersed in the electrolyte, and open circuit potential (OCP) was
allowed to stabilize for 15–30 minutes. Once a steady OCP was achieved, a polarization scan was initiated
from –250 mV below OCP to +1000 mV above OCP at a scan rate of 0.5–1.0 mV/sec. This scan captured the
anodic and cathodic responses of the material, allowing evaluation of passivation behaviour and pitting
susceptibility. The overlaid Tafel plots of all seven samples enabled a visual comparison of their corrosion
performance under identical testing conditions.

From the linear portions of the anodic and cathodic branches, Tafel extrapolation was used to determine key
corrosion parameters: the corrosion potential (Ecorr) and the corrosion current density (Icorr). These values
reflect the thermodynamic tendency and rate of corrosion, respectively. A lower Icorr indicates better corrosion
resistance, while a more positive Ecorr suggests greater passivation stability.

In addition to these electrochemical parameters, the corrosion rate was also calculated using the obtained Icorr
values. The corrosion rate (in mm/year) was determined by incorporating the density of the 17-4 PH stainless
steel, taken as 7.8 g/cm³, into the standard corrosion rate formula. This allowed for a quantitative estimation
of material degradation over time, providing further insight into the long-term performance of each heat-
treated condition in corrosive environments.

Fig. 3.10: Setup for corrosion resistance


testing

46
CHAPTER 4
RESULTS AND DISCUSSION

4.1 Microstructure Analysis

4.1.1 AR Sample
(a) (b)

Fig. 4.1.1: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Predominantly lath martensitic microstructure typical of 17-4 PH stainless steel in the solution-
annealed condition. The microstructure exhibits acicular features spread throughout the matrix, indicative of
martensitic transformation upon air cooling from high temperature. No visible evidence of precipitate
formation is observed at this stage, consistent with the absence of aging treatment.

47
4.1.2 SA Sample
(a) (b)

Fig. 4.1.2: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Refined and clearly defined lath martensitic microstructure, characteristic of 17-4 PH stainless
steel aged at 480 °C for 1 hour. The martensite laths are more pronounced compared to the as-received sample,
suggesting the initiation of Cu-rich precipitation within the matrix but the precipitates are too fine for optical
microscope.

4.1.3 DA1B Sample


(a) (b)

Fig. 4.1.3: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Indicates a tempered martensitic microstructure with more pronounced and coarsened lath
structures. The extended aging duration and higher secondary temperature have resulted in more significant
precipitation and possible partial coarsening of Cu-rich particles. The enhanced etching contrast and broader
laths suggest the presence of stress relief.

48
4.1.4 DA2B Sample

Fig. 4.1.4: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Tempered martensitic matrix with more pronounced etching and coarser lath boundaries
compared to DA1B. The elevated secondary aging temperature of 570 °C has resulted in further coarsening of
Cu-rich precipitates and possible grain boundary precipitation.

4.1.5 CA1B Sample

Fig. 4.1.5: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Fine, well-defined martensitic laths with sharp etching, indicating a refined and homogeneously
tempered structure. The cryogenic treatment prior to aging has transformed much of the retained austenite,
contributing to a higher volume fraction of martensite.

49
4.1.6 CA2B Sample

Fig. 4.1.6: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Finely refined martensitic laths with uniform spacing and moderate contrast, indicative of a
highly transformed martensitic matrix with minimal retained austenite. The extended immersion has further
increased dislocation density, providing abundant nucleation sites for Cu-rich precipitates during aging at
480 °C.

4.1.7 CA3B Sample

Fig. 4.1.7: (a): Mag.: 100X; (b): Mag.: 400X (etchant: Kaling’s Reagent 2)

Observation: Fine, well-defined martensitic laths, comparable to CA2B, but with minor thickening in
localized regions. Extended immersion has effectively eliminated retained austenite while introducing partial
recovery of dislocations. Subsequent aging at 480 °C produces fine Cu-rich precipitates that are slightly more
coarsened than those in CA2B.

50
4.2 Rockwell Hardness Results
Hardness peaks in CA2B (~48.7 HRC) and is lowest in DA2B (~38 HRC). SA (46.3 HRC) and CA1B/CA3B
(45.7 HRC) are relatively hard, whereas AR (40 HRC), DA1B (41.3 HRC) and especially DA2B (38 HRC)
are softer. These values follow expected trends: the lowest aging temperatures (e.g. H900) yield the hardest,
strongest condition, while higher aging (H1150, double-aging) softens the alloy. For example, Rolled Alloys
reports Rockwell-C hardness of ~44 at H900 and ~31 at H1150D. The CA treatments (especially CA2B)
evidently created the hardest microstructure, likely due to rapid precipitation of fine Cu clusters and little
tempering of martensite. Conversely, DA2B’s low hardness and high toughness suggest extensive precipitate
coarsening or martensite recovery. The inverse correlation between hardness and ductility/toughness (noted
below) underscores the classic trade-off in precipitation-hardening steels.

AVG. MICROHARDNESS
600
506.03
500 455.92
424.72 439.02
382.57
400
343.86
319.66
HV

300

200

100

0
AR SA DA1B DA2B CA1B CA2B CA3B

Fig. 4.2: Hardness (HRC) of 17-4PH samples: SA and CA2B treatments produced the
highest hardness values, whereas DA2B gave the lowest.

4.3 Microhardness Result Analysis

It is observed that different heat treatment conditions significantly influence the hardness values of 17-4 PH
stainless steel. The lowest hardness was observed for AR (343.86 HV), indicating a softer structure in the
untreated condition. SA showed an improved hardness value of 424.72 HV, which confirms that solution
annealing and aging enhance hardness through precipitation strengthening. Among all, CA2B demonstrated
the highest microhardness at 506.03 HV, followed by CA3B (455.92 HV) and CA1B (439.02 HV), indicating
that the combination of cryogenic treatment and controlled aging leads to a more refined and harder
microstructure. DA1B and DA2B showed relatively lower hardness values of 382.57 HV and 319.66 HV
51
respectively, suggesting that these conditions may have led to overaging or coarsening of precipitates, reducing
the hardness.

In conclusion, the trend in hardness clearly highlights that the CA2B condition is the most effective among
the tested heat treatment parameters for improving the hardness of 17-4 PH stainless steel. The improvement
can be attributed to enhanced martensitic transformation and uniform dispersion of strengthening precipitates
due to cryogenic treatment followed by optimal aging. The lower hardness in DA2B suggests that the double
aging cycle may require further optimization. Overall, the CA2B treatment is most favorable when increased
surface hardness and wear resistance are desirable.

AVG. MICROHARDNESS
600
506.03
500 455.92
424.72 439.02
382.57
400
343.86
319.66
HV

300

200

100

0
AR SA DA1B DA2B CA1B CA2B CA3B

Fig. 4.3: Microhardness (HV) of 17-4PH samples: CA2B and CA3B treatments produced
the highest HV values, whereas DA2B gave the lowest.

4.4 Charpy Impact Energy Analysis


Impact energy is highest for DA2B (~60.97 J) and DA1B (~53.56 J) and lowest for CA2B (~33.0 J). This
follows the classic inverse relationship with hardness: the most tempered (ductile) samples absorb more energy
before fracture, whereas the hardest, strongest conditions are more brittle. In 17-4PH, low aging (H900) gives
very low Charpy values (e.g. ~15 ft·lb, ~20 J), while overaged H1150-D can exceed ~100 ft·lb. Our DA results,
with moderate hardness and high impact energy, correspond to higher-aging treatments that promote
toughness. Conversely, the hard CA treatments have limited impact toughness, reflecting a martensitic
structure rich in fine precipitates. The observed correlation (toughness ∝ ductility and ∝ 1/hardness) is well-
documented: enhancing toughness by tempering usually costs tensile strength.

52
AVG. IMPACT ENERGY
70
60.966
60
53.56
50
Impact Energy (J)

38.3 39.733
40 35.38 36.033
33
30

20

10

0
AR SA DA1B DA2B CA1B CA2B CA3B

Fig. 4.4: Impact Energy (J) of 17-4PH samples: DA1B and DA2B treatments produced
the highest impact values, whereas CA2B gave the lowest.

4.5 Analysis Of YS, UTS and %Elongation

4.5.1 Yield Strength


The solution-annealed (SA) condition exhibits the highest yield force (~812 MPa), while the lowest yield force
occurs in the CA1B treatment (~481 MPa). In general, lower-temperature aging yields higher yield strength:
aging at ~482 °C (900 °F) produces peak hardness and strength. The SA and AR conditions (likely reflecting
extensive precipitation hardening) show the greatest stiffness, whereas the CA-treated samples have weaker
yield. This behavior aligns with 17-4PH metallurgy: fine, coherent Cu-rich precipitates formed at lower aging
temperatures (e.g. H900) raise yield strength, while higher-temperature or prolonged aging coarsens
precipitates and reduces yield. The observed yield trend (SA > AR > DA1B > CA3B > DA2B > CA2B >
CA1B) suggests SA achieved the most effective precipitation hardening, whereas CA1B produced a softer
martensitic matrix.

53
AVG. YIELD STRENGTH
900
812
800
710 685
700 653.5
617
600 568.5
YS (MPa) 481
500
400
300
200
100
0
AR SA DA1B DA2B CA1B CA2B CA3B

Fig. 4.5.1: Yield Strength (MPa) of 17-4PH samples: AR and SA


treatments produced the highest values, whereas CA1B gave the lowest.

4.5.2 Ultimate Tensile Strength


UTS rises in a similar pattern: CA1B has the highest tensile load (2101 MPa), followed by CA3B and CA2B,
whereas SA has the lowest (1276.5 MPa). In 17-4PH, maximum UTS is achieved by tempering at relatively
low temperatures (e.g. ~482 °C) but with sufficient time for precipitate growth. Our data indicate the CA
conditions yielded the greatest UTS, suggesting these treatments promoted optimal precipitation
strengthening. By contrast, SA’s low UTS could imply insufficient or uneven precipitation. The general trend
(CA1B ≈ CA3B ≈ CA2B > DA2B > DA1B > AR > SA) reflects known behavior: as aging temperature/time
increase, UTS first rises (due to fine precipitates) and then falls if overaging occurs. The high UTS of CA1B
correlates with its high hardness, consistent with literature: e.g. 17-4PH tempered at 900–925 °F attains UTS
≈ 1200–1300 MPa (≈174–189 ksi).

AVG. ULTIMATE TENSILE STRENGTH


2500
2101
1931.5 1969
2000 1790.3
1591.6
1508
UTS (MPa)

1500 1276.5

1000

500

0
AR SA DA1B DA2B CA1B CA2B CA3B

Fig. 4.5.2: UTS (MPa) of 17-4PH samples: CA1B and CA3B treatments produced the
highest values, whereas SA gave the lowest

54
4.5.3 %Elongation
Ductility (elongation) is greatest in DA2B (~19.80%) and DA1B (~19.0%), and least in CA3B (~15.45%).
The pattern (DA2B ≈ DA1B ≈ CA2B > AR ≈ CA1B > SA ≈ CA3B) matches expectations: softer, overaged
conditions (higher temper) yield higher elongation. For example, Rolled Alloys data show elongation climbing
from ~14% at H900 to ~22% at H1150-M. Here, DA treatments clearly enhanced ductility (at some expense
of strength), indicating more tempering or austenite reversion. In contrast, CA3B’s low elongation and high
hardness suggest a brittle martensite with minimal tempering. These trends are consistent with a precipitation-
hardening mechanism: low-temperature aging locks in hardness with low plasticity, while higher-temperature
aging or double-aging allows matrix relaxation and thus greater strain capability.

AVG. %ELONGATION
25

19.801
20 18.997 18.836
17.3 17.04 17.158
15.4545
%Elongation

15

10

0
AR SA DA1B DA2B CA1B CA2B CA3B

Fig. 4.5.3: %Elongation of 17-4PH samples: DA1B and DA2B treatments produced the
highest values, whereas CA3B gave the lowest

4.5.4 Stress-Strain Curve

From the stress-strain graph, it is evident that higher Yield Strength (YS) corresponds to reduced formability,
as the material resists deformation more strongly from the outset. The As-Received and CA1B samples, with
lower YS, exhibit better ductility and thus higher formability compared to CA2B or DA1B which have higher
YS but limited elongation.

Regarding toughness, which relates to the area under the stress-strain curve, DA1B and DA2B show superior
toughness due to a balance of high strength and good elongation. CA2B, although strong, has a steeper rise
and early failure, indicating less energy absorption before fracture. SA displays moderate toughness due to
lower strength but high ductility.

55
In conclusion, DA2B offers the best combination of strength and toughness, while CA2B excels in strength
but compromises ductility. Lower YS samples like AR and CA1B show better formability. Optimal
performance requires balancing strength with sufficient elongation for real-world applications.

Fig. 4.5.4: Superimposed Stress-Strain Curves with YS and UTS markers

56
4.6 XRD Analysis

Table 4.6.1: XRD Peak Shift and Phase Analysis Table for as received sample

Sample Reference Plane Shift Cause of the Shift


Peak (2θ) Peak (2θ) (hkl) Direction
~44.0° 44.5° (110) Negative Residual tensile stress or retained austenite causing
lattice expansion
~65.0° 65.2° (200) Negative Slight lattice expansion from microstrain or defects
~82.0° 82.3° (211) Negative Lattice distortion, likely from retained austenite or
processing-induced stress

X-ray diffraction (XRD) analysis was performed on the as-received 17-4 PH stainless steel to evaluate the
phase constitution and detect any shifts in lattice parameters. The diffractogram revealed three prominent
peaks located at approximately 44°, 65°, and 82°, corresponding to the (110), (200), and (211) planes of a
BCC structure, respectively. These peaks match well with the standard JCPDS reference pattern for martensitic
17-4 PH stainless steel.

A closer comparison with JCPDS reference data showed slight negative shifts (towards lower 2θ values) in all
57
three peaks. This downward shift typically indicates lattice expansion, which can result from:

• Residual tensile stresses are introduced during mechanical processing such as rolling, machining, or
welding.
• The presence of retained austenite, which has a larger lattice parameter than martensite and can slightly
alter the overall diffraction pattern.
• Microstructural defects or dislocations, contributing to local distortions and peak broadening or
shifting.

58
Table 4.6.2: XRD Peak Shift and Phase Analysis Table for standard aging sample

Sample Reference Peak Shift


Plane (hkl) Cause of the Shift / Observation
Peak (2θ) (2θ) Direction

Formation of retained or reversed austenite after


~40.2° 43.5° (γ(111)) (111) – γ Negative
aging

Slight lattice expansion, possibly due to residual


~43.8° 44.5° (110) – α Negative
stress or aging effects

Possible retained austenite or Cu-rich precipitate


~57° 59.0° (γ/χ(220)) (220) – γ/χ Negative
contribution

Slight
~66° 65.2° (200) – α Minor lattice contraction, stress relief from aging
Positive

~82° 82.3° (211) – α Aligned No significant shift

The XRD pattern of the standard aging sample (SA) reveals significant changes compared to the as-received
condition. Peaks corresponding to the martensitic (α) BCC structure are still dominant; however, the
emergence of additional peaks at ~40.2° and ~57° indicates the presence of retained or reversed austenite
(γ) and/or precipitate phases formed during the aging process.
Notably, the (111) γ peak at ~40.2° is a clear indication of partial reversion of martensite to austenite, a
phenomenon commonly observed during aging treatments at intermediate temperatures. The shift of the main
(110) α peak from 44.5° to ~43.8° suggests slight lattice expansion, possibly due to residual tensile stress or
the incorporation of interstitial atoms.

The standard aging treatment (solution annealing at 1040 °C followed by aging at 480 °C) promotes the
precipitation of Cu-rich or NiAl particles, which may also contribute to peak broadening or shifts.
Additionally, stress relaxation due to the thermal cycle can result in minor peak shifts toward standard
positions, as seen for the (200) and (211) α peaks.

Overall, the appearance of austenite peaks alongside shifted martensitic peaks confirms microstructural
evolution during the aging treatment, crucial for tailoring mechanical properties such as strength and
toughness in 17-4 PH stainless steel.

59
Table 4.6.3: XRD Peak Shift and Phase Analysis Table for DA1B sample

Sample Reference Peak Plane Shift Cause of the Shift / Observation


Peak (2θ) (2θ) (hkl) Directio
n

~40.1° 43.5° (γ(111)) (111) Negative Retained/reverted austenite formation; stabilized by


–γ thermal exposure

~43.9° 44.5° (110) Negative Slight lattice expansion due to Cu precipitation and
–α residual tensile stress

~65.2° 65.2° (200) None Peak matches reference; stable martensite phase
–α

~82.3° 82.3° (211) None Aligned with reference; stress-relieved and well-aged
–α martensite

The XRD pattern for the double aging sample (DA1B) shows dominant peaks at ~43.9°, 65.2°, and 82.3°,
corresponding to the (110), (200), and (211) planes of the BCC martensitic structure. An additional peak at
~40.1° indicates the presence of retained or reversed austenite (γ phase), which is more pronounced compared
to the standard aged condition.

The double aging process (480 °C for 2 h + 550 °C for 4 h) results in a more refined microstructure with well-
60
developed Cu-rich precipitates and partial reversion of martensite to austenite. The peak at ~43.9° shows
a slight negative shift compared to the JCPDS value (44.5°), likely due to lattice expansion from precipitate
formation and residual tensile stress.

The peak at ~50.4° is not a match for primary martensitic peaks and may correspond to ε-Cu precipitates,
which are known to form during aging in 17-4 PH SS.

Peaks at 65.2° and 82.3° closely align with standard reference values, indicating that the martensitic phase is
well-stabilized and the double aging process has relieved much of the internal stress introduced during
prior processing steps.

The presence of austenite along with stabilized martensite reflects a balanced microstructure, expected to
yield improved mechanical properties such as enhanced toughness and strength due to precipitation
hardening and microstructural refinement.

61
Table 4.6.4: XRD Peak Shift and Phase Analysis Table for DA2B sample

Sample Peak Reference Peak Plane Shift


Cause of the Shift / Observation
(2θ) (2θ) (hkl) Direction

Increased retained/reverted austenite due to higher aging


~40.0° 43.5° (γ(111)) (111) – γ Negative
temperatures

Slight lattice expansion from thermal exposure or Cu/NiAl


~43.95° 44.5° (110) – α Negative
precipitates

Likely due to Cu-rich precipitate (ε-Cu phase) or fine


~50.2° - - -
carbide/nitride

Slight Near-aligned; indicates stress relaxation and more


~65.3° 65.2° (200) – α
Positive complete precipitation

~82.3° 82.3° (211) – α None Peak matches reference; stable martensite

The XRD pattern of the Double Aging 2 sample (DA2B) demonstrates dominant BCC martensite peaks at
~43.95°, 65.3°, and 82.3°, corresponding to the (110), (200), and (211) planes. A noticeable peak at ~40.0°
indicates the presence of reversed or retained austenite, more pronounced than in the standard aged or DA1B
samples.

The double aging treatment at higher temperatures (520 °C and 570 °C) promotes enhanced atomic diffusion,
leading to increased Cu-rich or NiAl precipitation and partial reversion of martensite to austenite. The peak
shift of the α(110) reflection from 44.5° to ~43.95° suggests lattice expansion, likely due to internal stress
changes or solute redistribution.

The near-perfect alignment of (200) and (211) peaks with reference values implies a more relaxed and
thermodynamically stable martensitic structure, attributed to the longer and hotter aging sequence.

Compared to DA1B, the (110) peak at 43.6° is slightly closer to the reference 44.7°, indicating slightly
reduced lattice strain or smaller Cu clusters compared to DA1B.
The presence of the ~50.6° peak again suggests precipitation of ε-Cu, which is known to form more
prominently in double aging treatments.

62
Table 4.6.5: XRD Peak Shift and Phase Analysis Table for CA1B sample

Sample Reference Plan Shift Cause of the Shift / Observation


Peak (2θ) Peak (2θ) e Directi
(hkl) on

~44.05° 44.5° (110) Negativ Minor lattice expansion or residual stress from cryo-aging and
–α e precipitation effects

~65.2° 65.2° (200) None Perfect match; stable BCC martensitic structure
–α

~82.3° 82.3° (211) None Perfect match; no stress-induced distortion


–α

— 43.5° — — No austenite peak; cryo treatment helped suppress austenite,


(γ(111)) promoting martensite formation

The XRD pattern of the CA1B sample reveals dominant martensitic (BCC α) peaks at ~44.05°, 65.2°, and
82.3°, corresponding to the (110), (200), and (211) planes, respectively. Notably, no significant peak is
observed at ~40°, indicating an almost complete absence of retained or reversed austenite in the
microstructure.

The introduction of cryogenic treatment before aging plays a critical role in enhancing the martensitic phase

63
fraction. The deep sub-zero treatment (−80 °C to −196 °C) facilitates the transformation of residual
austenite into martensite, especially in a structure that has just undergone solution annealing. This
contributes to improved dimensional stability and hardness in the final aged condition.

The main (110) peak shows a slight shift toward lower 2θ (from 44.5° to 44.05°), which suggests minor lattice
expansion, possibly due to the introduction of dislocations or internal strain during the cryogenic cycle. The
aging treatment at 480 °C helps precipitate fine Cu-rich phases, enhancing strength while stabilizing the
martensite.

In summary, CA1B demonstrates a well-stabilized, fully martensitic structure with minimal austenite and
stress relaxation, making it a promising treatment route for applications demanding superior mechanical
performance and dimensional accuracy.

64
Table 4.6.6: XRD Peak Shift and Phase Analysis Table for CA2B sample

Sample Reference Plane Shift Cause of the Shift / Observation


Peak (2θ) Peak (2θ) (hkl) Directio
n

~44.08° 44.5° (110) – α Negative Minor lattice expansion due to high internal stress or
dislocations introduced by cryo soaking

~65.2° 65.2° (200) – α None Matches JCPDS; confirms BCC martensitic phase
stability

~82.4° 82.3° (211) – α Slight Minor contraction likely due to high degree of strain
Positive relaxation during cryo-aging + aging

— 43.5° — — No γ-phase detected; confirms nearly full martensitic


(γ(111)) transformation

The XRD pattern of the CA2B sample shows well-defined BCC martensitic peaks at approximately 44.08°,
65.2°, and 82.4°, corresponding to the (110), (200), and (211) planes, respectively. No peak is observed near
43.5°, indicating the absence of retained or reverted austenite, affirming the effectiveness of the extended
cryogenic treatment.

Compared to standard JCPDS data, the main (110) peak is slightly shifted to lower 2θ, indicating lattice
expansion, possibly due to the formation of dislocations or interstitial strain introduced during the prolonged
exposure to cryogenic temperatures (−196 °C for 6 hours). The minor shift in the (211) peak toward higher 2θ
may suggest local relaxation of residual stresses or precipitation-induced lattice contraction.

Peak broadening and secondary peaks (~50.5°, 88.2°) suggest increased Cu precipitation (ε-phase) and/or
secondary phase formation during the more aggressive heat treatment.

Overall, the CA2B condition demonstrates an advanced microstructural state with complete martensitic
transformation and negligible retained austenite, attributed to the synergistic effect of extended cryogenic
soaking and aging.

65
Table 4.6.7: XRD Peak Shift and Phase Analysis Table for CA3B sample

Sample Reference Plan Shift Interpretation


Peak (2θ) Peak (2θ) e Directio
(hkl) n

~44.2° 44.5° (110) Negativ Slight lattice expansion, possibly due to residual compressive
–α e stresses after cryogenic soak

~65.3° 65.2° (200) Slight Minor shift could relate to stress relaxation or interstitial
–α Positive redistribution

~82.5° 82.3° (211) Slight Similar behavior to (200); indicates refined martensitic matrix
–α Positive

— 43.5° — — No retained austenite observed


(γ(111))

The XRD profile of CA3B samples exhibits distinct peaks at ~44.2°, 65.3°, and 82.5°, which correspond to
the (110), (200), and (211) planes of the BCC martensitic phase. These closely match the standard 17-4 PH
JCPDS reference for martensite, with no evidence of austenitic γ(111) reflections, indicating a completely
martensitic structure.

66
The slight negative shift in the (110) peak compared to the reference may indicate residual compressive
stresses or lattice distortion induced during the extended cryogenic exposure (total of 10 hours, with 9 hours
in LN₂). The marginal positive shifts of (200) and (211) peaks suggest stress relaxation and possibly minor
precipitate-induced contraction due to the subsequent aging at 480 °C.

The presence of secondary peaks at ~50.5° and ~88.1° supports the formation of precipitates or complex
carbide/nitride phases due to overaging from triple-stage aging after cryogenic treatment.

Compared to CA1B and CA2B, the CA3B treatment shows peak positions converging toward standard
martensite references, indicating a well-stabilized, refined martensitic matrix after extensive cryogenic and
aging processes.

4.7 Corrosion Resistance Analysis


A clear trend can be observed regarding the corrosion resistance of 17-4 PH stainless steel under different heat
treatment conditions. The as-received (AR) sample demonstrates the lowest corrosion current density (Icorr)
and the lowest corrosion rate (0.145 mm/year), indicating the best corrosion resistance among all conditions.
In contrast, the cryogenically treated sample CA3B exhibits the highest Icorr value of 623.7 nA and the highest
corrosion rate of 4.031 mm/year, reflecting the poorest corrosion resistance. This trend suggests that as the
heat treatment complexity increases, particularly with double aging and cryogenic treatments, the corrosion
resistance tends to decrease due to changes in microstructure, such as increased precipitation or residual stress,
which may influence electrochemical activity.
In conclusion, the standard aging (SA) and cryogenic treatment group CA1B display moderate corrosion rates,
indicating some compromise between strength-enhancing treatments and corrosion performance. Notably, the
double aging and prolonged cryogenic treatments (DA1B, DA2B, CA2B, CA3B) lead to significantly higher
corrosion rates, suggesting that such treatments, while potentially improving mechanical properties, may
adversely affect the passive film stability or increase susceptibility to localized corrosion. Therefore, in
applications where corrosion resistance is critical, careful consideration must be given to selecting appropriate
heat treatment parameters, balancing mechanical enhancement with electrochemical stability.

67
Table 4.7: Tafel Extrapolation

PARAMETER Icorr (nA) Ecorr (mV) Rate(mm/yr)


AR 18.66 -557.4 0.145
SA 178.2 -519.6 1.042
DA1B 494.3 -567.6 3.036
DA2B 392.8 -585.5 2.838
CA1B 62.76 -551.6 0.385
CA2B 196.2 -585 1.338
CA3B 623.7 -587.4 4.031

Fig. 4.7: Overlay of Tafel Extrapolation of AR, SA, DA1B, DA2B, CA1B, CA2B, CA3B samples

68
CHAPTER 5
CONCLUSION

The present study successfully demonstrates how different heat treatment strategies—standard aging, double
aging, and cryogenic aging—influence the microstructure, mechanical performance, phase constitution,
and corrosion behavior of 17-4 PH stainless steel. By correlating microstructural evolution with test data, the
following key insights emerge:

5.1 Microstructural Findings:

• As-Received (AR) samples showed fine lath martensite with retained austenite and no visible
precipitates.

• Standard Aging (SA) promoted Cu-rich nanoprecipitate formation, enhancing strength and hardness
while refining laths.

• Double Aging (DA1B/DA2B) led to broader, tempered laths and precipitate coarsening, balancing
strength and toughness but risking over-aging.

• Cryogenic Aging (CA1B–CA3B) effectively transformed retained austenite and created dense, fine
precipitate dispersion. CA2B provided the most refined and stable martensitic matrix.

5.2 Mechanical Property Summary:

• SA yielded the highest Yield Strength (0.812 kN) due to optimal precipitation hardening.

• CA1B achieved the highest Ultimate Tensile Strength (2.101 kN); CA2B showed the highest
hardness (48.66 HRC).

• DA2B provided the best ductility (19.8%) and impact toughness (60.97 J), ideal for shock-sensitive
applications.

• CA3B showed improved strength but lower ductility, suggesting overexposure to cryogenic conditions
leads to brittleness.

69
5.3 XRD and Phase Analysis:

• AR exhibited both martensite and retained austenite.

• All treated samples displayed fully martensitic structures.

• Cryogenic and double aging treatments caused slight peak shifts and broadening, indicating
precipitation, residual stress relief, and lattice distortion.

• CA2B and CA3B showed sharp, stable α peaks, confirming a well-refined and low-stress martensitic
matrix.

5.4 Corrosion Resistance (Tafel Test):

• AR condition had the lowest corrosion rate (0.145 mm/year), showing best resistance.

• CA1B performed well with a rate of 0.385 mm/year, balancing transformation and passivity.

• Double Aging (DA1B/DA2B) and CA3B showed poor corrosion resistance due to increased
precipitates and internal stresses (CA3B worst at 4.031 mm/year).

• SA and CA2B offered moderate corrosion resistance.

5.5 Overall Conclusion:

• SA is optimal when high yield strength and balanced properties are required.

• DA2B is recommended for high ductility and impact toughness, making it suitable for dynamically
loaded parts.

• CA2B is ideal for applications needing maximum hardness and strength, such as wear-resistant
tools.

• AR offers the best corrosion resistance, suitable for chemically aggressive environments.

5.6 Final Verdict:

CA2B provides the best balance of strength and hardness, while DA2B is superior in toughness and
ductility. SA serves as a reliable baseline for strength-focused applications with moderate corrosion
performance.

70
Recommendation by Application:

Application Type Recommended Treatment


High strength/wear CA2B
High toughness/shock DA2B
Balanced strength/corrosion SA
Corrosion-sensitive parts AR

71
CHAPTER 6
FUTURE SCOPE

1. Fracture Surface Analysis of Tensile tested sample under Scanning Electron Microscope(SEM) and EDS.

72
APPENDICES

Appendix A: Hardness Of 17-4 PH SS

Load: 150kgf
Hardness: Rockwell Hardness (HRC)

1. AR:
Sr. no. Lapped surface Oxide layer
1 40 42
2 41 39
3 39 39
Average 40 40

2. SA:
Sr. no. Lapped surface Oxide layer
1 46 49
2 48 46
3 45 52
Average 46.33 49

3. DA1B:
Sr. no. Lapped surface Oxide layer
1 42 39
2 41 40
3 41 40
Average 41.33 39.66

4. DA2B:
Sr. no. Lapped surface Oxide layer
1 37 34
2 38 41
3 39 36
Average 38 37

5. CA1B:
Sr. no. Lapped surface Oxide layer
1 42 44
2 49 43
3 46 44
Average 45.66 43.66

73
6. CA2B:
Sr. no. Lapped surface Oxide layer
1 48 50
2 48 46
3 50 47
Average 48.66 47.66

7. CA3B:
Sr. no. Lapped surface Oxide layer
1 45 44
2 47 42
3 47 46
Average 45.66 44

Appendix B: Microhardness of 17-4 PH SS


Load: 1000gf
Dwell time: 15 seconds
1. AR:
Sr. no. D1 D2 HV
1 72.25 72.29 355.06
2 71.61 71.66 361.37
3 76.68 76.73 315.16
Average 73.51 73.56 343.86

2. SA:
Sr. no. D1 D2 HV
1 66.54 66.59 418.53
2 62.74 62.78 470.8
3 69.71 69.12 384.84
Average 66.33 66.16 424.72

3. DA1B:
Sr. no. D1 D2 HV
1 68.44 69.76 388.38
2 67.81 71.03 384.83
3 72.24 68.49 374.52
Average 69.49 69.76 382.57

74
4. DA2B:
Sr. no. D1 D2 HV
1 79.21 75.47 310.02
2 76.68 74.2 325.85
3 76.68 74.83 323.13
Average 77.52 74.83 319.66

5. CA1B:
Sr. no. D1 D2 HV
1 65.91 60.88 461.44
2 62.74 62.78 470.8
3 69.71 69.12 384.84
Average 66.12 64.26 439.02

6. CA2B:
Sr. no. D1 D2 HV
1 60.84 61.51 495.51
2 60.2 60.88 505.9
3 61.47 58.34 516.7
Average 60.83 60.24 506.03

7. CA3B:
Sr. no. D1 D2 HV
1 65.91 65.32 430.75
2 62.74 62.15 475.59
3 63.37 63.42 461.43
Average 64 63.63 455.92

Appendix C: YS, UTS, %Elongation Of 17-4 PH SS


1. AR:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 515 1942 3.846
2 976 1024 19.231
3 641 1559 28.846
Average 710 1508 17.30

2. SA:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 963 1039 7.273
2 661 1514 27.778
3 0 NAN 16.071
Average 812 1276.5 17.04

75
3. DA1B:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 516 1939 24.074
2 992 1008 13.559
3 547 1828 19.298
Average 685 1591.6 18.997

4. DA2B:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 481 2079 13.559
2 455 2199 28.302
3 915 1093 17.544
Average 617 1790.3 19.801

5. CA1B:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 515 1940 24.074
2 514 1946 13.115
3 414 2417 14.286
Average 481 2101 17.158

6. CA2B:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 399 2508 22.222
2 0 NAN 20
3 738 1355 14.286
Average 568.5 1931.5 18.836

7. CA3B:
Sr. no. YS (MPa) UTS (MPa) %Elongation
1 - 2750 -
2 769 1300 12.727
3 538 1858 18.182
Average 653.5 1969 15.4545

Appendix D: Charpy Impact Test Of 17-4 PH SS


1. AR:
Sr. no. Impact (J)
1 24.5
2 47.1
3 35.9
Average 35.83

76
2. SA:
Sr. no. Impact (J)
1 NA
2 41.8
3 34.8
Average 38.3

3. DA1B:
Sr. no. Impact (J)
1 49.2
2 50.8
3 60.7
Average 53.56

4. DA2B:
Sr. no. Impact (J)
1 72.1
2 50.6
3 60.2
Average 60.966

5. CA1B:
Sr. no. Impact (J)
1 42.3
2 30.3
3 35.5
Average 36.033

6. CA2B:
Sr. no. Impact (J)
1 22
2 29
3 48
Average 33

7. CA3B:
Sr. no. Impact (J)
1 31.8
2 29.6
3 57.8
Average 39.733

77
REFRENCES

1. J. Charles, "The history and development of stainless steels," Revue de Métallurgie, vol. 106, no. 2, pp.
104–113, 2009.

2. G. Krauss, Steels: Processing, Structure, and Performance, 2nd ed., ASM International, 2015.

3. M. J. Cieslak et al., “Aging behavior of 17-4PH stainless steel,” Metallurgical Transactions A, vol. 19, no.
9, pp. 2319–2331, Sep. 1988.

4. P. F. Thomason, Precipitation Hardening, Butterworths, 1990.

5. AK Steel, “17-4 PH Stainless Steel – Product Data Sheet,” [Online]. Available: https://www.aksteel.com

6. M. Escriba, J. Llanes, and E. Martín, “Influence of aging treatments on the precipitation hardening behavior
of 17-4PH stainless steel,” Materials Characterization, vol. 60, no. 5, pp. 404–410, 2009.

7. M. Liu et al., "Microstructure evolution and mechanical properties of 17-4 PH SS under various cryogenic
and double aging treatments," Materials Science and Engineering A, vol. 726, pp. 205–213, 2018.

8. "Stainless Steel - Grade 17-4 (UNS S17400)," AZoM. [Online]. Available:


https://www.azom.com/article.aspx?ArticleID=6778rolledmetalproducts.com+2AZoM+2Wikipedia+2

9. "The Role Chromium Plays in Stainless Steel," CIVMATS. [Online]. Available:


https://www.civmats.com/news/news46/news46.HTMLcivmats.com

10. "Nickel Institute," Nickel Institute. [Online]. Available: https://nickelinstitute.org/media/2309/niadv-


en.pdfNickel Institute+1Reddit+1

11. "Sequential nucleation of phases in a 17-4PH steel: Microstructural evolution and mechanical properties,"
ScienceDirect. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S1359645416309193ScienceDirect

12. "The role of niobium in austenitic and duplex stainless steels," Niobium Tech. [Online]. Available:
https://niobium.tech/-/media/niobiumtech/attachments-biblioteca-tecnica/nt_the-role-of-niobium-in-
austenitic-and-duplex-stainless-steels.pdfNiobium+1Niobium+1

13. "ATI 17-4™ Technical Data Sheet," ATI Materials. [Online]. Available:
https://www.atimaterials.com/Products/Documents/datasheets/stainless-specialty-
steel/precipitationhardening/ati_17-4_tds_en_v2.pdfatimaterials.com

14. "17-4 Precipitation-Hardening Stainless Steel: Soft-Tough Heat Treatment," PMC. [Online]. Available:
https://pmc.ncbi.nlm.nih.gov/articles/PMC11642631/PMC+1Alro Steel+1

15. "Tailoring microstructure and mechanical properties of 17-4PH steel via heat treatment," ScienceDirect.
[Online]. Available:
https://www.sciencedirect.com/science/article/abs/pii/S2214860423005675ScienceDirect

16. "Formability, Microstructure and Mechanical Properties of Flow Formed 17-4 PH Stainless Steel,"
Springer. [Online]. Available: https://link.springer.com/article/10.1007/s11665-018-3724-9SpringerLink
78
17. ATI 17-4™ Technical Data Sheet, ATI Materials. [Online]. Available:
https://www.atimaterials.com/Products/Documents/datasheets/stainless-specialty-
steel/precipitationhardening/ati_17-4_tds_en_v2.pdfatimaterials.com

18. "Effect of heat treatment on microstructure and mechanical properties of 17-4PH stainless steel,"
ScienceDirect. [Online]. Available:
https://www.sciencedirect.com/science/article/pii/S2238785423020951ScienceDirect

19. "Data Sheet 17-4 Stainless Steel - AMS 5643 - UNS S17400," SSA Corp. [Online]. Available:
https://www.ssa-corp.com/documents/Data%20Sheet%2017-4-Stainless-Steel-AMS-5643-
UNSS17400.pdfssa-corp.com

20. "Data Sheet - 17-4 Stainless - Rolled Alloys," Rolled Alloys. [Online]. Available:
https://www.rolledalloys.com/wp-content/uploads/17-4_Data-sheet-rolled-alloys.pdfRolled Alloys

21. "ARMCO® 17-4 PH® - AK Steel," AK Steel. [Online]. Available:


https://www.aksteel.nl/files/downloads/clf_datasheet_armco_17-4_ph_pdb_euro_102022_89.pdfaksteel.nl

22. "Influence of aging treatments on 17–4 PH stainless steel parts," SpringerLink. [Online]. Available:
https://link.springer.com/article/10.1007/s00170-023-11136-3SpringerLink

23. "17-4 PH/H900/H1025/H1075/H1150 | Vested Metals," Vested Metals. [Online]. Available:


https://www.vestedmetals.net/17-4/vestedmetals.net

24. "17-4PH Stainless Steel - Beartech Alloys," Beartech Alloys. [Online]. Available:
https://www.beartechalloys.com/17-4PH-stainless-steel.htmlbeartechalloys.com

25. "17-4PH Annealed - Penn Stainless," Penn Stainless. [Online]. Available:


https://www.pennstainless.com/resources/product-information/stainless-grades/precipitation-hardening-
grades/17-4ph-annealed/Penn Stainless+1Penn Stainless+1

26. "17-4PH H1025 - Penn Stainless," Penn Stainless. [Online]. Available:


https://www.pennstainless.com/resources/product-information/stainless-grades/precipitation-hardening-
grades/17-4ph-h1025/Penn Stainless

27. "Stainless Steel Alloy 17-4PH Precipitation Hardening," Sandmeyer Steel. [Online]. Available:
https://www.sandmeyersteel.com/alloy-17-4ph/Sandmeyer Steel Company

28. "Properties of 17-4 PH Stainless Steel Material - boyi technology," Boyi Prototyping. [Online]. Available:
https://www.boyiprototyping.com/materials-guide/17-4-stainless-steel/BOYI

29. "17-4ph Stainless Steel | sus 630 stainless steel | H900 | H1150," eSteel Suppliers. [Online]. Available:
https://www.esteelsuppliers.com/new/shop/precipitation-stainless-steel/17-4ph-sus630-1-4548-stainless-
steel/E Steel Sdn.Bhd

30. "CUSTOM 630 (17-4) - Carpenter Technology," Carpenter Technology. [Online]. Available:
https://www.carpentertechnology.com/hubfs/7407324/Material%20Saftey%20Data%20Sheets/Custom%206
30%20%2817-4%20PH%29.pdfCarpenter Technology

79
31. "Laser powder bed fusion of 17–4 PH stainless steel: A comparative study," ScienceDirect. [Online].
Available: https://www.sciencedirect.com/science/article/pii/S2214860421003390ScienceDirect
[18] "A Review of the Mechanical Properties of 17-4PH Stainless Steel Produced by Bound Powder
Extrusion," MDPI. [Online]. Available: https://www.mdpi.com/2504-4494/7/5/162MDPI

32. P.-Y. Hsieh et al., “17-4 PH Precipitation‑Hardening Stainless Steel: Soft-Tough Heat Treatment
Mechanism and Thermal Fatigue Characteristics,” Materials, vol. 17, no. 23, p. 5851, Nov. 2024.

33. M. Villa et al., “Aging 17-4 PH Martensitic Stainless Steel Prior to Hardening: Effects on Martensitic
Transformation, Microstructure and Properties,” Materialia, vol. 32, p. 101882, 2023.

34. A. G. Rakhshtadt, A. Yu. Akimova, and V. S. Arzamasova, “Double aging of stainless steels,” Metalloved.
i Term. Obrab. Metallov, vol. 13, no. 11, pp. 983–984, Nov. 1971.

35. P. Jurči and I. Dlouhý, “Cryogenic Treatment of Martensitic Steels: Microstructural Fundamentals and
Implications for Mechanical Properties and Wear and Corrosion Performance,” Materials, vol. 17, no. 3, p.
548, Jan. 2024.

80
ACKNOWLEDGEMENT

We express our deep feelings of gratitude to our guide Dr. S. U. DANGRIKAR , Professor
of Department of Metallurgy and Materials Science, College of Engineering, Pune, for the
valuable advice and guidance in the completion of this project work. We are greatly indebted
to him for instilling in us the idea and structure of this project report. We would like to
acknowledge our collaborators and colleagues, who have constantly helped, encouraged,
boosted our morale, and have worked with us in completing this dissertation work.

We are also thankful to Dr. Mukesh Kumar, Mr. Abhijit Bhopale, Ashish Sir and Shruti
Ma’am for his valuable assistance in carrying out experimental activities.

We are grateful to our parents for consistently supporting and motivating us.

Sincerely,

Aman Raju Kadu

Saloni Sameer Sakala

81

You might also like