You are on page 1of 22

Additive Manufacturing 58 (2022) 103050

Contents lists available at ScienceDirect

Additive Manufacturing
journal homepage: www.elsevier.com/locate/addma

Review

Towards standardizing the preparation of test specimens made with


material extrusion: Review of current techniques for tensile testing
Clara Phillips a, Mark Kortschot b, Fae Azhari a, *
a
Department of Mechanical and Industrial Engineering, University of Toronto, 5 King’s College Rd., Toronto, ON M5S 3G8, Canada
b
Department of Chemical Engineering and Applied Chemistry, University of Toronto, 200 College St., Toronto, ON M5S 3E5, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: Researchers and industries rely heavily on standardized testing methods to ensure products are designed and
Material extrusion manufactured with high quality. With increasing use of additive manufacturing processes, such as material
Additive manufacturing extrusion (MEX), there is a significant gap in material testing standards tailored for such components, and a lack
Testing standards
of guidance on preparing appropriate test specimens with mesostructures that best represent the final part being
Material characterization
analyzed. This paper aims to support the standardization of MEX material testing by reviewing the current
Tensile testing
Slicing process parameters methods used for preparing test specimens for tensile testing and proposing guidelines for implementation in a
new standard. The need for standardization of MEX specimen preparation is addressed by analyzing the effects of
slicing parameters on resulting tensile properties of the specimen. It is suggested that a standard should
acknowledge these parameters, in addition to specimen geometry, toolpath optimization, printer and material
specifications, so that they are appropriately selected for the test specimen by regarding the final part structure.
Consideration of the proposed guidelines in a standardized method may enable comparisons between published
results and support the development of MEX technology for use in advanced applications.

1. Introduction materials such as steel, cement and wood [3], and by the 1920s, many
manufacturers required their material suppliers to adhere to ASTM
1.1. History of testing standardization standards. Substantial advances in test methods, inspection procedures
and production techniques were made during the war efforts [4] and
Standardized systems support modern civilization, connecting the with newly found chemistries of synthetic materials in the 1920s and
globe in the form of communication, supply-chain infrastructure, and 1930s [5].
regulated manufacturing. In its early incarnations, standardization was Polymeric materials originally fell under the jurisdiction of Com­
developed to reach a common system of alphabet characters, units of mittee D09 on electrical insulators, but as the use of polymers rapidly
measurement, and currency. As civilization evolved, standardization expanded to other applications in the 1940s, Committee D20 on plastics
expanded to facilitate international connections, disseminate knowledge was formed [5]. Complementary standards developed by the ASTM
from scientific and technological innovations, and enable successful Committee E28 on Mechanical Testing in 1969 undoubtedly improved
quality assurance and control, leading to products and services that are the rigor of technical testing procedures already under the jurisdiction of
safe, better trusted by users, and supported by authority [1]. material-specific committees like the D20. A proliferation of more
The American Society for Testing and Materials (now ASTM Inter­ detailed and specific testing procedures followed. The widely used
national) was established in 1898 and comprises over 130 technical ASTM D638 for tensile testing of plastics was first documented in 1984
committees of experts working to reach consensus on standards for [6]. Until the development of ASTM D3039 in 1976 [7] by Committee
various testing procedures and materials [2]. It currently serves as the D30, polymer composite industries were either directed to D20 testing
American representative of the International Organization for Stan­ standards for plastics, or resorted to developing their own proprietary
dardization (ISO) sector of testing and materials, which was founded standards to meet their needs [8]. This is not uncommon, as scientific
later in 1947 [2]. Early ASTM Committees focussed on specifications of advances usually supersede the slower, more bureaucratic task of

* Corresponding author.
E-mail address: azhari@mie.utoronto.ca (F. Azhari).

https://doi.org/10.1016/j.addma.2022.103050
Received 21 March 2022; Received in revised form 1 July 2022; Accepted 18 July 2022
Available online 22 July 2022
2214-8604/© 2022 Elsevier B.V. All rights reserved.
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

standardization [9]. Table 1


Today, industries face a vast selection of testing methods to deter­ Additive manufacturing standardized procedures.
mine material properties. ISO has been adopted by the EU and most Standard Name Intended AM
European countries. However, the more detailed ASTM International category
standards are used globally across many industries, because ISO test ISO 17296–2:2015 Additive manufacturing – General principles G
procedures are limited [10]. In many instances, ASTM and ISO standards – Part 2: Overview of process categories and
are written for the same testing objectives, but differ in technical pro­ raw material
cedures; therefore, test data produced by the two procedures are not ISO 17296–3:2014 Additive manufacturing – General principles G
– Part 3: Main characteristics and
directly comparable [10]. corresponding test methods
When selecting the appropriate standard to use, the degree to which ISO 17296–4:2014 Additive manufacturing – General principles G
procedures can be modified to meet testing objectives should also be – Part 4: Overview of data exchange
considered. Where the focus is to compare test results with other labo­ ISO/ASTM Additive manufacturing – General principles G
52900:2021 – Fundamentals and vocabulary
ratories, following exact protocols is crucial. Sometimes, though, it is
ISO/ASTM Additive manufacturing – General principles G
necessary to veer from standard methods in order to obtain important 52901:2017 – Requirements for purchased AM parts
material characteristics that help define or predict basic mechanisms of ISO/ASTM Additive manufacturing – Test artifacts – G
a component in its intended end-use application. If this is done, test 52902:2019 Geometric capability assessment of additive
documentation must capture all sample preparation and procedure manufacturing systems
ISO/ASTM Additive manufacturing – Material MEX
techniques used and include reasonable justifications for any deviations. 52903–1:2020 extrusion-based additive manufacturing of
The fundamental assumptions and scientific principles on which stan­ plastic materials – Part 1: Feedstock
dard procedures are originally based must be understood to appropri­ materials
ately interpret test results from a modified procedure. ISO/ASTM Additive manufacturing – Material MEX
52903–2:2020 extrusion-based additive manufacturing of
plastic materials – Part 2: Process equipment
1.2. Additive manufacturing and standardization ISO/ASTM Additive Manufacturing – Process PBF
52904:2019 characteristics and performance – Practice
As additive manufacturing (AM) progresses from rapid prototyping for metal powder bed fusion process to meet
critical applications
to producing load-bearing componentry [11,12], the demand for stan­
ISO/ASTM Additive manufacturing – Feedstock PBF
dardized mechanical testing of AM parts increases in parallel. AM is 52907:2019 material – Methods to characterize metal
already starting to transform the aerospace industry [13], the automo­ powders
tive sector [14], sporting wear and helmets [15], and biomedical ap­ ISO/ASTM Additive manufacturing – Design – G
plications of prostheses, dental implants, bio-printed tissues, surgical 52910:2018 Requirements, guidelines and
recommendations
tools and more [16]. However, to truly revolutionize these industries
ISO/ASTM Additive manufacturing – Design – Part 1: PBF
with AM and to ensure safe designs, our understanding of the structural 52911–1:2019 Laser-based powder bed fusion of metals
properties of AM components must reach the same level as that of ISO/ASTM Additive manufacturing – Design – Part 2: PBF
conventionally made components. In the medical device industry, bench 52911–2:2019 Laser-based powder bed fusion of polymers
testing is a major avenue for creating scientific evidence for regulatory ISO/ASTM Specification for additive manufacturing file G
52915:2020 format (AMF) Version 1.2
decision-making [17]. New aerospace components undergo rigorous ISO/ASTM Standard terminology for additive G
material, proof-of-concept, and component tests before compliance for 52921:2013 manufacturing – Coordinate systems and
certification is met, and only at this stage may major tests with the test methodologies
aircraft be performed [18]. ISO/ASTM Additive manufacturing – System PBF
52941:2020 performance and reliability – Acceptance
The most popular AM method is material extrusion (MEX) [19,20],
tests for laser metal powder-bed fusion
commonly referred to as fused filament fabrication, or fused deposition machines for metallic materials for
modeling. In MEX, filament material is heated and dispensed through a aerospace application
nozzle, as defined in ISO/ASTM 52900 [21]. Other common AM ASTM F2971–13 Standard Practice for Reporting Data for G
methods are vat photopolymerization, in which a liquid photopolymer is Test Specimens Prepared by Additive
Manufacturing
selectively cured, and powder bed fusion (PBF), in which regions of a ASTM F3091–14 Standard Specification for Powder Bed PBF
powder bed are selectively fused with thermal energy [21]. PBF is the Fusion of Plastic Materials
most widely used process for functional industrial-use parts [22]. ASTM F3122–14 Standard Guide for Evaluating Mechanical G
However, fueled by the maker community and small companies, MEX Properties of Metal Materials Made via
Additive Manufacturing Process
remains the most popular method for tinkering, prototyping and
ASTM F3335–20 Standard Guide for Assessing the Removal of PBF
building esthetic pieces due to its low upfront and consumable costs Additive Manufacturing Residues in Medical
[23], and large material availability [24]. Yet, MEX is limited in use for Devices Fabricated by Powder Bed Fusion
functional parts, as structural properties of MEX parts are generally
G – General AM (i.e., not specific to an AM method)
subpar compared to those produced by PBF [25]. MEX – Material extrusion AM
Standardized test methods for AM primarily focus on PBF technol­ PBF – Powder bed fusion AM
ogies (Table 1). PBF is a highly controlled method due to its use of
proprietary software and materials, and thus enables consistency by
the production of functional parts beyond the prototyping context [13,
reducing human input [26,27]. Standardization of stringent PBF pro­
16,33]. Standards for powder-based AM methods were likely partly
tocols is more easily attained compared to MEX, which is generally
fueled by published peer-reviewed articles, such as Seifi et al. [34] and
considered to be an open-sourced method due to its greater flexibility in
Gorelik [35] in 2017, that summarized the limitations of standards
using third-party materials and manipulating process parameters [28,
available at the time and provided technical considerations for future
29]. Furthermore, the superior mechanical properties resulting from
standards. It is hoped that a comprehensive review for MEX technologies
PBF renders it a better candidate for fabricating critical load-bearing
will, likewise, accelerate MEX standardization development.
parts that require regulation and testing, such as in aerospace and
medical device industries. Nevertheless, as MEX technologies and ma­
terials continue to advance (e.g., by incorporating carbon fiber rein­
forcement and other additives [30–32]), MEX is expanding to include

2
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

1.3. Testing specimens representing a final part proposal of practical guidelines for specimen preparation to propel the
development of a new AM standard.
Advancement of MEX technology requires accurate quantification of
material properties through detailed standards tailored to MEX speci­ 3. State of AM standardized testing
mens [36]. The first step in predicting the structural behavior of an MEX
part is material characterization through tensile testing of representative 3.1. ASTM and ISO standards
coupon specimens. However, tensile test standards for MEX parts are
non-existent, hence there is limited guidance for preparing appropriate Tensile tests are performed on material samples according to ASTM
specimens. Numerous research groups have investigated the tensile D638 or ISO 527–2 for polymers, or ASTM D3039 [65] for polymer
properties of MEX specimens, as reviewed in [37–42]. However, these composites. Although there are fundamental differences between ISO
studies mainly follow ASTM D638 [43] or ISO 527–2 [44] standards, 527–2 and ASTM D638 methods for polymers, agreement in results may
which have been used inconsistently [45–47] and have limited appli­ be found, depending on the material used [66].
cability for MEX specimens [48–50]. Section 3.2 discusses the applica­ At the time of this publication, no standard explicitly addresses
bility of current standards. tensile testing of AM polymer specimens. Garcia-Dominguez et al. [50]
The properties of the polymer comprising the filament feedstock may and Forster [48] provide a detailed compilation of the current state of
be measured with single filament specimens [51–54], with 100% infill ISO and ASTM standards applied to the AM field. Both articles conclude
3D printed specimens with surface notches removed [46,55], or with that the current state of standards is incomplete for the material char­
conventional injection molded specimens [56,57]. In contrast, the acterization of AM specimens. In Table 1, we provide a non-exhaustive
properties of a printed portion of a load-bearing structure depend on list of these AM standards.
both the polymer filament and the complex mesostructure created by the Evidently, the standardization industry is looking ahead. In 2011,
slicer software and printing parameters chosen. This mesostructure in­ ISO mobilized a Technical Committee, “ISO/TC 261 Additive
cludes such things as the infill percentage, the raster pattern, and the manufacturing”, and added formal cooperation agreements in 2020
perimeter contours. An MEX part is a classic hierarchical structure, [67]. Their forward-thinking is also demonstrated with the established
where the structure at a variety of size scales influences the properties. standard ISO/ASTM 52915 for a digital additive manufacturing file
Having chosen a base polymer filament, however, it is the mesostructure format. This was developed on the basis that AM is evolving quickly
which can vary widely in a load-bearing part. It is not practical to model from homogeneous parts, to those requiring a file format that can cap­
all the details of the mesostructure when performing a finite element ture more complex features such as color, multi-material settings, gra­
analysis of a structural component, such as a load-bearing beam or shell, dients, and microstructures.
and in this case, the designer must use equivalent homogenized prop­ ISO 17296–3:2014 [68] and ISO/ASTM 52921:2013 [69] provide
erties of the as-printed “material” [58–64]. In order to do this, appro­ the most relevant guidance for mechanical testing of AM specimens, but
priate and representative test specimens must be fabricated. It is point to non-AM standards for direction. For example, ISO
obviously not enough to simply design a tensile specimen to be printed 17296–3:2014 refers to ISO 527–2 [44] for tensile testing, and ISO/­
and tested with arbitrary slicing software settings. A testing standard ASTM 52921:2013 [69] refers to both ISO 527–2 [44] and ASTM D638
will have to provide explicit guidance about how to ensure that the [43] for tensile testing. In 2015, the National Institute of Standards and
mesostructure of the printed tensile specimen represents the final part Testing analyzed the applicability of non-AM ASTM and ISO standards
being analyzed. for AM specimens, and concluded that ISO 527–2 and ASTM D638 may
While the material science community has studied the relation be­ provide a good approximation of the results but should be used with
tween the mesostructure and mechanical properties quite extensively, caution until AM-based standardization is further developed [48].
there are no testing standards making use of this knowledge. In this
review, the most important aspects of the extruded structure are dis­ 3.2. Standardized test specimen preparation
cussed to serve as the basis for the development of an effective stan­
dardized testing method useful to engineers designing load-bearing AM Aspects of ASTM D638, ISO 527–2 and ASTM D3039 such as testing
parts. set up and procedure, strain rate, and grip specifications have significant
effects on resulting tensile properties [10], and can feasibly be applied to
2. Objectives AM specimens. However, problems occur when it comes to specimen
preparation from feedstock material, which is the focus of this paper.
This review, organized as follows, is intended to provide clarity Conventional materials are molded or casted, with a limited number of
about the current preparation methods used for MEX tensile test speci­ process parameters affecting material properties (e.g., molding tem­
mens, and ultimately contribute to new standardization protocols perature, injection pressure, flow rate and flow direction [70]). MEX,
tailored for MEX: however, is not only affected by these same properties for raw filament
material, but must also deal with numerous additional parameters in the
1. A review of relevant ASTM and ISO testing standards. creation of the GCode (through slicing software or direct GCode gen­
2. Identification of consistencies and discrepancies in test specimen eration) prior to fabrication.
preparation methods used in previous research on tensile properties There are three main reasons for caution when adapting ASTM D638,
of MEX specimens. ISO 527–2 or ASTM D3039 for MEX specimens:
3. Evaluation of proposed guidelines for MEX specimen preparation, to
be considered for implementation in a new AM standard. 1. These standards do not warn nor indicate that the material properties
of AM specimens are significantly influenced by unavoidable and
Although this review focuses on coupon specimen preparation for often unpredictable void formation between layers, by the thermal
tensile testing, the proposed guidelines may be translated to specimen history from the re-melt and solidification cycle of each layer deposit
preparation for bending, fatigue, impact, and other tests. This review is [20,71], and by slicing settings such as build orientation (Fig. 1A),
intended to be a continuation of previous rigorous reviews on the cur­ raster angle (Fig. 1B), layer height, raster width, air gap, number of
rent state of ASTM and ISO standards and their applicability for AM, contours (Fig. 1C), and printing temperature and speed [49,50,72,
particularly those by Dizon et al. [39], Forster [48], and 73]. In its introduction, ISO 17296–3 briefly addresses this point, but
Garcia-Dominguez et al. [49,50]. The unique contribution of this review still refers to ISO 527 as a recommended test standard, demon­
is its detailed focus on specimen preparation techniques, and its strating the need for updated standards and caution when using

3
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Fig. 1. MEX process parameters. (A) View of three coupons fabricated in the three major build orientations, (B) rendered section views of filament rasters laid in
three different raster patterns for each of the three major build orientations subjected to different tensile loading directions, and (C) detailed view of raster
cross-sections.

current recommendations. The reader must be aware of these effects perimeter contours. There is also limited understanding of the effect
and select the suitable parameters by being mindful of the final part of directly fabricating the desired geometry versus machining the
structure. geometry from a blank, which imposes new stresses and strains at the
2. These standards recommend preparing test specimens by injection perimeter of the specimen. This is discussed in Section 5.1. Ulti­
molding, compression molding, or machining from materials in mately, to obtain meaningful results, the extruded material grain size
casted or extruded sheets, plates, slabs or similar forms [43,44]. should match that of the representative part. In MEX parts, filament
However, as shown in Section 4, studies on tensile properties of MEX rasters are held together by small contact forces with neighboring
materials mainly fabricate their test specimens by extruding directly rasters. Machining may compress these deposited rasters into one
to the desired geometry. There are a few exceptions [56,74], which another, altering the specimen properties and surface roughness,
are discussed in Sections 4.6 and 5.1. Direct extrusion of ASTM D638 thus better representing a hybrid (subtractive and additive) manu­
or ISO 527–2 specimen geometry can result in either at least one factured part. A comparison of the perimeter of a machined specimen
perimeter contour around the test coupon, or numerous notches on to a specimen printed with one contour is shown in Fig. 2.
the perimeter surface if no contour is used (Figs. 1B and 1C). This has 3. Visual inspections prescribed in ISO 527–1 Section 6.4 are imprac­
a significant impact on material properties [36, 75–77], with the tical for MEX specimens. The layer-by-layer deposition unavoidably
general consensus that tensile strength increases with the number of results in mesoscale surface notches (Fig. 1C), thus specimens would

Fig. 2. Test specimens machined from an MEX plate (left), and test specimens fabricated directly to size (right).

4
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

need to be rejected according to the visual inspection criteria set by process parameters on tensile properties. MEX process parameters are
ISO 527–1. Furthermore, as demonstrated by Allum et al. [78], generally categorized in literature based on (1) build, (2) material, and
measuring an accurate load-bearing cross-sectional area of the (3) environmental and printer specifications. A fourth category, test
specimen may require microscopic imaging due to the presence of specimen specifications, is added here to address specimen preparation
surface notches and voids. Reliance on current standard techniques, considerations that are not related specifically to MEX, but that are
like simply measuring specimen thickness and width with digital typically included in a standard specimen preparation method. These
calipers, would overestimate the cross-sectional area of an MEX process and preparation parameters are provided in Table 2.
specimen. A new standard must be introduced to guide MEX spec­ Research reviews by [37–42] have aimed to understand the effects of
imen inspections and measurements. process parameters on final part mechanical properties. Table 3 sum­
marizes these reviews by providing an extensive, but non-exhaustive,
Consistency and repeatability in test specimen preparation are still list of publications that have investigated the tensile properties of
challenging as the complex AM processes are yet to be completely MEX specimens, and identifies, at a minimum, the following parameters
characterized and understood. In a comparative study, Miller et al. [79] used in each publication: build orientation, material and MEX printer
concluded that ASTM D3039 specimens more consistently adhered to used, referenced standard for geometry of the test specimens, layer
failure acceptance criteria (i.e., failing within the specimen’s gauge height, raster angle, raster width, infill percentage, air gap, number of
length), and resulted in higher elastic moduli but lower tensile strengths perimeter contours, fabrication method, and number of samples per test.
compared to ASTM D638 specimens. ASTM D3039 specimens are rect­ These parameters were selected due to their frequent use in the litera­
angular whereas ASTM D638 specimens are dog-bone shaped with a ture as either constant parameters, or as testing conditions. Any addi­
width that tapers toward the center. The induced stress concentrations tional parameters that were explicitly treated as testing conditions are
at this change in width are amplified in certain build orientations by the also included in Table 3. Additional studies [46,47,78,81], which were
staircase effect from layer deposition, perhaps explaining the inconsis­ not mentioned in reviews [37–42], are included in Table 3 since they
tent ultimate strength and failure location in the tapered region. The provide an analysis of fabrication methods that are directly relevant to
study goes on to suggest using more samples than the required five when this review. Although other material property results were reported in
using ASTM D638 to achieve at least five sample failures inside the some of these publications (e.g., fatigue strength or hardness), for
gauge region [79]. These geometries are discussed in further detail in simplicity, only the main conclusions for ultimate tensile strength (UTS)
Section 5.1. and modulus (where reported) are presented.
The only applicability of ASTM D638, ISO 527–2 and ASTM D3039 Most of the peer-reviewed literature has reached a consensus on the
standards for AM may be in their recognition of the build orientation general effects of printing parameters on UTS and modulus; however,
parameter. These standards recommend that, for potentially anisotropic direct comparisons between the various results reported cannot be made
materials, two types of specimens should be prepared: one with their due to vast differences in methods used (as is evident in Table 3 and
long axes parallel with, and the other normal to, the principal axes of further discussed in Section 6). The general trends are discussed in the
anisotropy. ISO 17296–3 also requires documenting the test direction following sections, and exceptions are highlighted.
and build orientation in the test report. As AM knowledge grows, more
process parameters must be included in standard test reports.
Creating standard procedures is challenging, particularly for 4.1. Build orientation
advancing fields like AM, where technologies and materials are
continually evolving [71,80]. The standardization development must Build orientation was investigated in 14 of the 37 studies reviewed.
account for the many degrees of freedom in design that is enabled by Two studies looked only at XY and YZ [75,99]; four at XY and Z [77,86,
AM’s layer-by-layer production technique. Fully characterizing AM 96,107]; one at YZ and Z [78]; six at XY, YZ and Z [87,88,101,104,105,
processes and how various variables affect mechanical properties re­ 108]; and one study did not clearly describe tested orientations [109]. In
quires significant time and large amounts of data. Research reviews, general, YZ specimens were strongest and stiffest, followed by XY, then
such as this one, are vital to highlight consistencies and discrepancies Z. Greater loads and less deformation is realized when loading is par­
that help direct future standardization. allel, rather than transverse, to the filament lines, explaining why
significantly worse performance is found in Z specimens. For example,
3.3. Terminology UTS from YZ to Z specimens decreased by 56 % in [88] and by 23 %
[105]. In YZ specimens, the cross-sectional area is narrow, making the
As the objective of this review is to support the development of an
AM standard, we use the standardized terminology outlined in ISO/ Table 2
ASTM 52900 wherever possible. However, inconsistent terminology and MEX process and preparation parameters.
limited visuals used in the literature for build orientation have made it 1. Build Specifications • Build orientation
difficult to assign an accurate three-letter orthogonal orientation nota­ • Layer height
Raster angle
tion for the reviewed references listed in Table 3. Therefore, we use the •
• Raster width
notation for bilateral symmetry to describe the three major orientations, • Infill percentage
such that XY denotes any flat orientation of a coupon, YZ denotes an • Infill pattern
orientation of a coupon on its side, and Z denotes a vertical orientation • Air gap
(Fig. 1A). • Number of perimeter contours
2. Material Specifications • Material composition
• Color
4. Tensile characterization of MEX specimens • Density
3. Environmental and Printer Specifications • 3D printer type
Current literature on tensile characterization of MEX specimens uses • Nozzle diameter
Nozzle temperature
a large variety of MEX process parameters, and many publications often •
• Build envelope temperature
do not clearly report on the specifications used. To determine reasonable • Extrusion speed
parameter settings for recommendation by a standard, we must first 4. Test Specimen Specifications • Referenced standard
fully understand what effect these parameters have on tensile test re­ • Test specimen dimensions
sults, and which parameters have the greatest effect. Therefore, in this • Fabrication method
Number of samples
section, we summarize the current understandings of the effects of MEX

5
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3
Review of publications investigating tensile properties of MEX specimens.
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Lanzotti et al. Build XY Layer height 0.1, 0.12, 0.15, 0.18, 0.2 0.15 mm layer height, 0˚raster angle and 4
[36] orientation: (mm): contours resulted in the highest UTS of
Material: PLA (brand not specified) Raster angle: 0˚, 18˚, 45˚, 72˚, 90˚ 53.6 MPa (SD: 0.77 MPa).
3D printer: Prusa i3 Contours: 2, 3, 4, 5, 6 0.1 mm layer height, 0˚raster angle and 3
Standard: Modified ASTM D638 Type contours resulted in the highest modulus of
II 3.50 GPa (SD: 0.14 GPa).
Raster width –
(mm):
Infill: 100 %
Air gap: Outline overlap of 15 %
Fabrication: Extruded to size
# Samples 3
per test:
Striemann et al. Build Z Layer height 0.2, 0.3, 0.4 0.2 mm layer height and polishing resulted
[46] orientation: (mm): in the highest UTS of 9.0 MPa (SD: 0.4 MPa).
Material: CarbonX™ Nylon, Fabrication: Extruded to size with no polishing; extruded 0.2 mm layer height and no polishing
3DXTECH (carbon fiber- to size with stepwise polishing (grain sizes of resulted in the highest modulus of 1.68 GPa
reinforced polyamide) 180, 320, 600 then 1000) (SD: 0.026 GPa).
3D printer: Ultimaker 2 + Extended
Standard: DIN 53442
Raster angle: 0˚
Raster width 0.5
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: 1
# Samples 3 (minimum)
per test:
Zhang et al.[47] Build XY Standard: ASTM D638 Type I and II; ISO 527–2 Type ISO 20753 A12 and extruded to size resulted
orientation: 1 A; ISO 20753 A12; ASTM D3039 in the highest UTS of approximately 71 MPa
Material: Stratasys ULTEM 9085 Fabrication: Extruded to size; water jet cut from extruded (SD bars shown, approximately 5 MPa) and
plate highest modulus of approximately 2.6 GPa
3D printer: Stratasys Fortus 450 MC (SD not reported).
Layer height 0.254
(mm):
Raster angle: ±45◦
Raster width 0.508
(mm):
Infill: 100 %
Air gap 0
(mm):
Contours: 1
# Samples 5
per test:
Croccolo et al. Material: Stratasys ABS-M30 Build XY, YZ XY orientation and 10 contours resulted in
[75] orientation: the highest UTS of 29.7 MPa (SD: 0.4 MPa),
3D printer: – Contours: 1, 4, 7, 10 and highest modulus of 2.12 GPa (SD:
Standard: Modified ASTM D638 Type 0.05 GPa).
II
Layer height 0.25
(mm):
Raster angle: ±45˚
Raster width 0.5
(mm):
Infill: 100 %
Air gap 0
(mm):
Fabrication: Extruded to size
# Samples 5
per test:
Cwikla et al. Build XY Infill: 10 %, 25 %, 40 %, 60 % 40% infill, 4 contours, honeycomb pattern,
[76] orientation: and 1x flow multiplier resulted in the highest
Material: ABS (brand not specified) Contours: 1, 2, 4, 7 UTS of approximately 42.5 MPa. SDs not
3D printer: Prusa i3 Flow rate 0.5, 0.75, 1 reported.
multiplier: Modulus not reported.
Standard: ISO 527 1BA Infill pattern: Honeycomb, rectilinear concentric, top
concentric
Layer height –
(mm):
Raster angle: –
Raster width –
(mm):
(continued on next page)

6
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Air gap –
(mm):
Fabrication: Extruded to size
# Samples 5
per test:
Torrado and Material: Cycolac grade MG47 ABS Build XY, Z XY orientation, ASTM D638 Type IV
Roberson[77] orientation: standard geometry and 0˚raster angle
3D printer: Custom Lulzbot TAZ 4 Standard: ASTM D638, Type I, IV, V resulted in the highest UTS of 39.9 MPa (SD:
Layer height 0.27 Raster angle: 0˚, 0˚/90˚ 1.4 MPa).
(mm): Modulus not analyzed, but results reported
Raster width – Contours: 0, 1 similar values across samples.
(mm):
Infill: 100 %
Air gap –
(mm):
Fabrication: Extruded to size
# Samples 5
per test:
Allum et al.[78] Material: 3DXTECH Natural PLA Build YZ, Z YZ orientation, 0.1 mm layer height and
orientation: 0.5 mm raster width resulted in the highest
3D printer: RepRap X400 Layer height 0.1, 0.15, 0.2, 0.25, 0.3 UTS of 68.7 MPa (SD bars shown,
(mm): approximately 5 MPa).
Standard: ASTM D1708 Raster width 0.4, 0.45, 0.5, 0.55, 0.6 Modulus not reported.
(mm):
Raster angle: 0˚
Infill: 100 %
Air gap –
(mm):
Contours: –
Fabrication: Razor blade cutting from
custom extruded boxes
# Samples 6
per test:
Miller et al.[79] Build XY 3D printer: Dimension, uPrint ASTM D638 Type IV standard geometry,
orientation: 0.254 mm layer height, ± 45˚raster angle,
Material: Stratasys ABS P430 Standard: ASTM D638 Type I, Type IV and ASTM and solid infill resulted in the highest UTS of
D3039 30.52 MPa (SD bars shown, approximately
Raster width – Layer height 0.254, 0.3302 1.5 MPa).
(mm): (mm): ASTM D3039 standard geometry, 0.254 mm
Air gap – Raster angle: ±45˚, 0˚/90˚ layer height, 0˚/90˚raster angle, and solid
(mm): infill resulted in the highest modulus of
Contours: 1 Infill: “Solid”, “Sparse high density” 1.72 GPa (SD bars shown, approximately
Fabrication: Extruded to size 0.1 GPa).
# Samples 5
per test:
Koch et al.[81] Build XY Contours: 0, 2 2 contours and extruded to size resulted in
orientation: the highest UTS of 31.3 MPa (SD bars shown,
Material: MatterHackers ABS Fabrication: Extruded to size; water jet cut from extruded approximately 5 MPa).
plate; smoothing with acetone vapor Modulus not reported.
3D printer: LulzBot TAZ 5
Standard: ASTM D638 Type I
Layer height 0.2
(mm):
Raster angle: 90◦
Raster width –
(mm):
Infill: 95 %
Air gap 0
(mm):
Contours: 2
# Samples 5
per test:
Ahn et al.[82] Build XY Raster angle: 0˚, ± 45˚, 0˚/90˚, 90˚ 0˚raster angle and − 0.0508 mm air gap
orientation: resulted in the highest UTS of approximately
Material: Stratasys ABS P400 Raster width 0.508, 1.00 21 MPa (SD bars shown, approximately
(mm): 2 MPa).
3D printer: Stratasys 1650 Air gap (mm): 0, − 0.0508 Modulus not reported.
Standard: ASTM D3039 Material color: Blue, white
Layer height 0.275
(mm):
Infill: 100 %
Contours: –
Fabrication: Extruded to size
2
(continued on next page)

7
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

# Samples
per test:
Rajpurohit and Build XY Layer height 0.10, 0.15, 0.20, 0.25, 0.30 0.6 mm raster width, 0.1 mm layer height
Dave[83] orientation: (mm): and 0˚raster angle resulted in the highest UTS
Material: PLA (brand not specified) Raster angle: 0˚, 30˚, 45˚, 60˚, 90˚ of 47.3 MPa (SD: 2.69 MPa).
3D printer: OMEGA Dual Extruder Raster width 0.4, 0.5, 0.6, 0.7 Modulus not numerically investigated, but
(mm): graphical results showed an increase in
Standard: ASTM D638 Type I modulus with 0.1 mm layer height and
Infill: 100 % 0˚raster angle.
Air gap –
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 2
per test:
Garg and Build XY Layer height 0.178, 0.254, 0.330 0.178 mm layer height resulted in the
Bhattacharya orientation: (mm): highest UTS of approximately 32.5 MPa for
[84] Material: Stratasys ABS P430 one sample. Means and SDs were not
3D printer: Stratasys Uprint SE Plus reported.
and Mojo Modulus not numerically investigated, but
Standard: Modified ASTM D638 Type graphical results showed an increase in
I modulus with 0.178 mm layer height.
Raster angle: 0˚/90˚
Raster width –
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: –
Fabrication: “Trimmed” to size
# Samples 3
per test:
Ziemian et al. Build XY Raster angle: 0˚, 45˚, ± 45˚, 90˚ 0˚raster angle resulted in the highest UTS of
[85] orientation: 25.5 MPa (SD: 0.73 MPa), and highest
Material: ABS (composition: 90–100 modulus of 0.99 GPa (SD: 0.02 GPa).
% ABS resin, 0–2 % mineral
oil, 0–2 % tallow, 0–2 %
wax)
3D printer: Stratasys Vantage-I
Standard: ASTM D3039
Layer height 0.1778
(mm):
Raster width 0.3048
(mm):
Infill: 100 %
Air gap 0
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 5–10
per test:
Onwubolu and Material: Stratasys ABS P400 Build XY, Z XY orientation, 0.127 mm layer height, 45˚
Rayegani[86] orientation: raster angle, 0.2032 mm raster width, and
3D printer: Stratasys 400mc Layer height 0.127, 0.3302 − 0.00254 mm air gap resulted in the
(mm): highest UTS of 34.6 MPa. No SDs reported.
Standard: Modified ASTM D638, Type Raster angle: 0˚, 45˚ Modulus not reported.
unknown
Infill: 100 % Raster width 0.2032, 0.5588
(mm):
Contours: – Air gap (mm): -0.00254, 0.5588
Fabrication: Extruded to size
# Samples 3
per test:
Durgun and Material: Stratasys ABS P430 Build XY, YZ, Z YZ orientation and 0˚raster angle resulted in
Ertan[87] orientation: the highest UTS of approximately 36 MPa
3D printer: Dimension 3D Printer Raster angle: 0˚, 30˚, 45˚, 60˚, 90˚ (SD bars shown, approximately 0.5 MPa).
Standard: ISO527 Type 1B Z orientation and 90˚raster angle resulted in
Layer height 0.254 highest modulus of approximately 2.5 GPa
(mm): (SD bars shown, approximately 0.5 GPa).
Raster width – Note: Different terminology convention for
(mm): raster angle and build orientation were used,
Infill: – thus best assumptions were made to apply
Air gap – terminology consistent with the other
(mm): studies.
(continued on next page)

8
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Contours: –
Fabrication: Extruded to size
# Samples 3
per test:
Riddick et al. Material: Stratasys ABS-M30 Build XY, YZ, Z YZ orientation and 0˚raster angle resulted in
[88] orientation: the highest UTS of 34.17 MPa (SD:
3D printer: Stratasys Fortus 400 Raster angle: 0˚, ± 45˚, 0˚/90˚, 90˚ 1.47 MPa), and highest modulus of 2.79 GPa
Standard: ASTM D638 Type I (SD: 0.05 GPa).
Layer height 0.127
(mm):
Raster width 0.2032
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 5
per test:
Hill and Haghi Build XY Raster angle: 0˚, 15˚,30˚, 45˚, 60˚, 75˚, 90˚ 0˚raster angle resulted in the highest UTS of
[89] orientation: 59.78 MPa and highest modulus of 2.08 GPa
Material: Stratasys ABS (SDs not reported).
3D printer: Stratasys (model not Note: Conventional 0˚orientation was
specified) referred to as 90˚in[89].
Standard: ASTM D638 Type I
Layer height 0.267
(mm):
Raster width 0.508
(mm):
Infill: 100 %
Air gap 0
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 6
per test:
Es-Said et al. Build XY Raster angle: 0˚, ± 45˚, 45˚, 90˚, 0˚/45˚ 0˚raster angle resulted in the highest UTS of
[90] orientation: 20.6 MPa. SDs not reported.
Material: Stratasys ABS P400 Modulus not reported.
3D printer: Stratasys 1650
Standard: ASTM D638 Type I
Layer height –
(mm):
Raster width –
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: –
Fabrication: Extruded to size
# Samples 3 (for 0˚, ± 45˚, 0˚/45˚), 4 (for
per test: 45˚), 5 (for 90˚)
Casavola et al. Build XY Material: ABS, PLA (brands not specified) 0˚raster angle resulted in the highest UTS of
[91] orientation: 50.23 MPa (SD: 0.77 MPa) and highest
3D printer: Prusa i3 Raster angle: 0˚, 45˚, 90˚ modulus of 3.12 GPa (SD: 0.03 GPa) for PLA;
Standard: ASTM D638 Type I and UTS of 26.11 MPa (SD: 2.37 MPa) and
Layer height 0.35 modulus of 1.79 GPa (SD: 0.58 GPa) for ABS.
(mm):
Raster width 0.7
(mm):
Infill: 100 %
Air gap 0
(mm):
Contours: 2
Fabrication: Extruded to size
# Samples 5
per test:
Ziemian et al. Build XY Raster angle: 0˚, ± 45˚, 45˚, 90˚ 0˚raster angle resulted in the highest UTS of
[92] orientation: 25.15 MPa (SD: 0.45 MPa) and highest
Material: Stratasys ABS P400 modulus of 1.49 GPa (SD: 0.05 GPa).
3D printer: Stratasys Vantage-I
Standard: ASTM D638 Type I
Layer height 0.1778
(mm):
(continued on next page)

9
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Raster width –
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 4
per test:
Wu et al.[93] Build XY Material: PEEK (Jilin University Special Plastic 0/90˚raster angle and 0.3 mm layer height
orientation: Engineering Research Co.), Stratasys ABS resulted in the highest UTS of 56.6 MPa for
P430 PEEK, and 27.1 MPa for ABS. SDs not
Standard: GB/T 16421 3D printer: PEEK with custom 3D printer, ABS with reported.
uPrint SE 3D printer Modulus not numerically investigated, but
Raster width – Layer height 0.2, 0.3, 0.4 graphical results showed a higher modulus of
(mm): (mm): PEEK than ABS.
Infill: 100 % Raster angle: 0˚/90˚, 30˚/60˚, ± 45˚
Air gap 0
(mm):
Contours: 2
Fabrication: Extruded to size
# Samples 5
per test:
Rankouhi et al. Build XY Layer height 0.2, 0.4 0˚raster angle and 0.2 mm layer height
[94] orientation: (mm): resulted in the highest UTS of 39.4 MPa (SD:
Material: ABS (brand not specified) Raster angle: 0˚, 45˚, 90˚ 0.3 MPa).
3D printer: Makerbot Replicator 2x 45˚raster angle and 0.2 mm layer height
Standard: ASTM D3039 resulted in the highest modulus of 2.23 GPa
Raster width – (SD: 0.07 GPa).
(mm):
Infill: 100 %
Air gap 0
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 4
per test:
Nomani et al. Build XY Layer height 0.2, 0.4, 0.6, 0.8 0.2 mm layer height and 9 mm sample
[95] orientation: (mm): thickness resulted in the highest UTS of
Material: Bilby3D ABS Standard: ASTM D638 Type IV thicknesses: 3, 6, 9 mm 45.3 MPa and highest modulus of 2.22 GPa.
3D printer: Lulzbot Mini SDs not reported.
Raster angle: ±45˚
Raster width –
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: 2
Fabrication: Extruded to size
# Samples 5
per test:
Sebert et al. Standard: ASTM D638 Type I for XY, Build XY, Z ABS, XY: Similar UTS for both personal
[96] modified Type I for Z orientation: (30 MPa) and professional (29 MPa)
Layer height 0.25 for professional Material: Stratasys ABSi, Stratasys Nylon 12 printers.
(mm): printer, 0.2 for personal Nylon, XY: Similar UTS for both personal
printer (41 MPa) and professional (38 MPa)
Raster angle: 0˚/90˚ 3D printer: Stratasys Fortus 400mc (professional), printers.
MakerGear M2 (personal) For the Z orientation: Nylon and the
Raster width – professional printer resulted in the highest
(mm): UTS of 36 MPa.
Infill: 100 % SDs not reported.
Air gap – Modulus not reported.
(mm):
Contours: 0
Fabrication: Extruded to size
# Samples “At least 3′′
per test:
Zhang et al.[97] Build XY Standard: ASTM D638 Type I and II, ASTM D3039 ASTM D638 Type II standard geometry
orientation: resulted in the highest UTS of approximately
Material: ULTEM® 9085 65 MPa (SD bars shown, approximately
3D printer: Stratasys Fortus 900mc 0.5 MPa).
Layer height 0.254 Modulus not reported.
(mm):
Raster angle: ±45˚
(continued on next page)

10
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Raster width 0.508


(mm):
Infill: 100 %
Air gap 0
(mm):
Contours: –
Fabrication: Extruded to size
# Samples 3
per test:
Ouhsti and Build XY Raster angle: 0˚, 30˚, 60˚ 30˚raster angle, 190◦ C printing temperature,
Soufiane[98] orientation: and 50 mm/s printing speed resulted in the
Material: PLA (brand not specified) Printing 190, 200, 210 highest UTS of 23.6 MPa. SDs not reported.
temperature 0˚raster angle, 190◦ C printing temperature,
(◦ C): and 30 mm/s printing speed resulted in the
3D printer: Wanhao Duplicator 4 S Printing speed 30, 50, 70 highest modulus of 3.94 GPa. SDs not
(mm/s): reported.
Standard: ISO 527–2 Type 1BA
Layer height 0.2
(mm):
Raster width –
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: –
Fabrication: Extruded to size
# Samples 3
per test:
Keleş et al.[99] Material: Stratasys ABS P430 Build XY, YZ YZ orientation, rectangular geometry with
orientation: no hole, and ± 45˚raster angle resulted in the
3D printer: Stratasys uPrint SE Standard: ASTM D638 Type I, and rectangular highest UTS of 27.4 MPa (SD: 0.5 MPa).
specimens, with no hole or hole diameters: Modulus not reported.
1, 2, 3, 4, 5 mm
Layer height 0.254 Raster angle: ±45˚and 0˚/90˚for XY, ±45˚for YZ
(mm):
Raster width –
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: –
Fabrication: Extruded to size
# Samples 28 – 35
per test:
Rodríguez et al. Build XY Raster angle: 0˚, 10˚, 30˚, 45˚, 60˚, 90˚ 0˚raster angle and − 0.0254 mm air gap
[100] orientation: resulted in the highest UTS of 24.4 MPa (SD
Material: Stratasys ABS P400 Air gap (mm): -0.0254, 0.0762 not reported) and highest modulus of
3D printer: Stratasys FDM1600 1.97 GPa (SD: 0.02 GPa).
Standard: ASTM D3039
Layer height 0.254
(mm):
Raster width 0.305
(mm):
Infill: 100 %
Contours: –
# Samples 3 (minimum)
per test:
Zaldivar et al. Material: ULTEM® 9085 Build XY, YZ, Z, YZ tilted 45˚, Z tilted 45˚ YZ orientation and 45˚raster angle resulted in
[101] orientation: the highest UTS of 71.03 MPa (SD: 2.6%) and
3D printer: Stratasys Fortus 400mc Raster angle: 0˚, 90˚for XY, 45˚for YZ, 0˚for YZ tilted, Z, Z highest modulus of 2.48 GPa (SD: 7.7%).
tilted
Standard: ASTM D638 Type I
Layer height 0.254
(mm):
Raster width –
(mm):
Infill: –
Air gap –
(mm):
Contours: –
Fabrication: Extruded to size
# Samples –
per test:
XY Material color: Natural, black, gray, blue, white
(continued on next page)

11
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Wittbrodt and Build Natural PLA resulted in the highest UTS of


Pearce[102] orientation: 57.16 MPa (SD: 0.35 MPa).
Material: Lulzbot PLA All colors had similar modulus of 2.78 GPa
3D printer: (SD: 0.35 GPa).
Lulzbot TAZ 4
Standard:
ASTM D638 Type I
Layer height (mm):

Raster angle:
0˚/90˚
Raster width (mm):

Infill: 100 %
Air gap (mm): –
Contours: –
Fabrication: Extruded to
size
# Samples per 10
test:
Tymrak et al. Build XY Material and MOST RepRap (Natural ABS, Clear PLA); Details on individual parameter
[103] orientation: 3D printer: Lulzbot Prusa Mendel RepRap (Natural ABS, combinations were not specified. Average of
Standard: ASTM D638 Type I Purple PLA, White PLA); Prusa Mendel tested specimens with 0.2 mm layer height
Raster width – RepRap (Black PLA); Original Mendel for PLA resulted in the highest UTS of
(mm): RepRap (Natural PLA) 60.4 MPa, and highest modulus of 3.48 GPa.
Infill: 100 % SDs not reported.
Air gap –
(mm):
Contours: – Layer height 0.2, 0.3, 0.4
(mm):
Fabrication: Extruded to size Raster angle: ±45˚, 0˚/90˚
# Samples 10
per test:
Uddin et al. Material: Z-ABS Build XY, YZ, Z YZ orientation, 0˚raster angle and 0.19 mm
[104] orientation: layer height resulted in the highest UTS of
3D printer: Zortax M200 Layer height 0.09, 0.19, 0.39 approximately 30 MPa. SDs not reported.
(mm): YZ orientation, 0˚raster angle and 0.09 mm
Standard: ASTM D638 Type I Raster angle: 0˚, 45˚, 90˚ layer height resulted in the highest modulus
Raster width – of approximately 1.5 GPa. SDs not reported.
(mm):
Infill: 100 %
Air gap –
(mm):
Contours: 1
Fabrication: Extruded to size
# Samples 3 (minimum)
per test:
Cantrell et al. Standard: ASTM D638 Type IV Build XY, YZ, Z YZ orientation, PC material with Fortus
[105] orientation: 360mc, and ± 45˚raster angle resulted in the
Infill: 100 % Material and ABS (brand not specified) for Ultimaker 2; highest UTS of 58.0 MPa (SD: 1.4 MPa) and
3D printer: PC (brand not specified) for Stratasys Fortus highest modulus of 2.0 GPa (SD: 0.03 GPa).
360mc
Air gap 0 Layer height 0.1 for Ultimaker 2; 0.254 for Fortus 360mc
(mm): (mm):
Contours: – Raster angle: ±45˚, 30˚/60˚, 15˚/75˚, 0˚/90˚for XY; ±45˚, 0˚/90˚
for YZ and Z
Fabrication: Extruded to size Raster width 0.4 for Ultimaker 0.2; 0.508 for Fortus
(mm): 360mc
# Samples 10
per test:
Deng et al. Build XY Layer height 0.2, 0.25, 0.3 0.2 mm layer height, 40% infill, 60 mm/s
[106] orientation: (mm): printing speed, and 370◦ C printing
Material: PEEK (Jilin University Infill: 20%, 40%, 60% temperature resulted in the highest UTS of
Special Plastic Engineering 40.0 MPa (SD: 4.4 MPa).
Research Co.) 0.2 mm layer height, 60% infill, 40 mm/s
3D printer: Custom FDM printer Printing speed 20, 40, 60 printing speed and 360◦ C printing
(mm/s): temperature resulted in the highest modulus
Standard: ISO 527–2 Type 1B Printing 350, 360, 370 of 576.5 MPa (SD: 49.9 MPa).
temperature
(◦ C):
Raster angle: –
Raster width –
(mm):

(continued on next page)

12
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Table 3 (continued )
Reference Coupon preparation Main conclusions

Constant parameters Tested parameters

Air gap
(mm):
Contours: –
Fabrication: Extruded to size
# Samples –
per test:
Laureto and Material: PLA (multiple brands, not Build XY, Z XY orientation, ASTM D638 Type IV
Pearce[107] all specified) orientation: standard geometry, 0.15 mm layer height,
3D printer: Athena Delya RepRap Standard: ASTM D638 Type I and IV for XY; Type IV 210◦ C printing temperature, and 1x flow
for Z multiplier resulted in the highest UTS of
Raster angle: – Layer height 0.06, 0.09, 0.1, 0.15, 0.18, 0.2, 0.25, 0.3, 0.5 61.6 MPa. SDs not reported.
(mm): Modulus not reported.
Raster width – Printing speed 30, 40, 50, 55, 60, 67, 70, 75, 80, 85, 90, 200
(mm): (mm/s):
Infill: 100 % Printing 175, 180, 182, 185, 190, 200, 205, 210, 215,
temperature 220, 230
(◦ C):
Air gap: Infill overlap 15 % Flow rate 0.85, 0.9, 1, 1.04, 1.05
multiplier:
Contours: Shell thickness of 0.5 mm
Fabrication: Extruded to size
# Samples 3
per test:
Hossain et al. Material: Stratasys PC Build XY, YZ, Z YZ orientation, ± 45˚raster angle, 0.483 mm
[108] orientation: raster width, − 0.013 mm air gap, and
3D printer: Stratasys Fortus 900mc Raster angle: ±45˚, 30˚/60˚, 0˚/90˚ 0.508 mm contour width resulted in the
Standard: ASTM D638 Type I Raster width 0.432, 0.457, 0.483, 0.508 highest UTS of 63.96 MPa (SD bars shown,
(mm)* : approximately 2 MPa).
Layer height 0.254 Air gap (mm) -0.013, 0 The same parameters, but with a 0 mm air
(mm): *: gap, resulted in the highest modulus of
Infill: 100 % Contour width 0.432, 0.508, 0.559 approximately 1.88 GPa (SD bars shown,
(mm)* : approximately 0.04 GPa).
Contours: 1 * These parameters were varied in the slicing software solely to
Fabrication: Extruded to size minimize void density, thus not all parameter combinations
# Samples 5 were under investigation.
per test:
Sood et al.[109] Material: Stratasys ABS P400 Build 0˚, 15˚, 30˚(unclear what axis the angles are 0˚orientation, 0.254 mm layer height, 60˚
orientation: relative to) raster angle, 0.4064 mm raster width, and
3D printer: Vantage SE Layer height 0.127, 0.178, 0.254 0.008 mm air gap resulted in the highest UTS
(mm): of 18.09 MPa. SDs not reported.
Standard: ISO 527:1966 Raster angle: 0˚, 30˚, 60˚ Modulus not reported.
Infill: 100 % Raster width 0.4064, 0.4564, 0.5064
(mm):
Contours: 1 Air gap (mm): 0, 0.004, 0.008
Fabrication: Extruded to size
# Samples 3
per test:

ratio of contour to infill pattern much greater compared with XY spec­ tended to result in superior mechanical performance (with the exception
imens. Since contours are deposited parallel to the tensile loading di­ of Wu et al. [93]). Keleş et al. [99] posit that the combination of ± 45˚
rection, this may account for the higher UTS compared with XY raster angle with YZ orientation specimens exhibits the greatest UTS, but
specimens, reported in all but one [75] of the studies. since they did not test other raster angles with the YZ orientation, the
superior results may have been due to build orientation.
4.2. Raster angle
4.3. Air gap
Raster angle is the most extensively examined parameter (24 of 37
studies). Of the 18 studies that included a longitudinal raster angle (0˚), The effect of air gap was investigated in five of 37 studies. A negative
most reported greatest UTS and modulus achieved with this setting. For air gap improves UTS [82,86,100,108] and modulus [67,71,85] since it
example, Casavola et al. [91] found a 74.3 % increase in UTS and 36 % increases bonding strength between adjacent rasters. However, Sood
increase in modulus with raster angles of 90˚to 0˚. A few exceptions were et al. [109], reasoned that smaller air gaps restrict heat dissipation and
found: Rankouhi et al. [94] demonstrated greater modulus with 45˚than may increase the amount of residual stress in the part. Furthermore,
0˚, Ouhsti et al. [98] showed greatest UTS with 30˚, Sood et al. [109] their scanning electron microscopy observations suggest that leaving a
showed greatest UTS with 60˚, and Onwubolu and Rayegani [86] re­ positive air gap may allow material to flow between adjacent layers,
ported that varying the raster angle from 0˚to 45˚had very little effect on which increases the area of bonded surfaces [109]. More research is
UTS. Sood et al. [109] speculated that the longer raster depositions needed to validate this claim, however.
associated with smaller raster angles may increase the probability of
stress accumulation along that direction, resulting in more distortion 4.4. Perimeter contours
and weaker bonding and thus reduced strengths. For the six studies that
did not test 0˚raster angle as a parameter, but rather investigated the The number of perimeter contours was investigated in five of 37
effect of alternating rasters (e.g., ± 45˚, 0˚/90˚ and/or 30˚/60˚), ± 45˚ papers. More contours were found to increase UTS [75,81] and modulus

13
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

[75], whereas Torrado and Roberson [77] found greater UTS but little reduced surface roughness. However, Lanzotti et al. [36] and Sood et al.
effect on modulus. Lanzotti et al. [36] and Cwikla et al. [76] both found [109] speculated that lower layer heights, under a certain value, may
a peak in UTS and modulus around four contours. Only one study actually decrease UTS due to the increased number of layers required,
examined zero contours [77]. Aside from Z orientation, contours are laid which results in more heating and cooling cycles, thus inducing residual
in the direction of tensile loading, explaining why incorporating con­ stresses and interlayer distortion. Of the studies that also reported on
tours in the test specimen increases the UTS and modulus. modulus, all demonstrated highest moduli with lower layer heights [36,
46,79,83,84,94,95,103,104,106].
4.5. Specimen geometry Layer height and raster width may have large interactive effects with
other printing parameters as they both impact raster-to-raster bonding
Seven studies examined the effect of specimen geometry [47,77,79, and void density. They also affect the temperature gradient throughout
95,97,99,107]. Zhang et al. [97] found the highest UTS with ASTM the part since cooling rates vary with different raster volumes. Layer
D638 Type II geometry, attributing it to the smaller cross-sectional area, height and raster width also determine the flow rate from the nozzle,
and thus a greater contour-to-raster ratio, which may have increased the thus influencing the strain placed on the extruded material depending
UTS. This explanation was also provided by Miller et al. [79], who found on the extruder size, extruder temperature and extruder speed. As
increased tensile strength with geometries of smaller cross-section (Type further discussed in Section 4.9, more research is needed to examine
IV, compared to Type I and ASTM D3039). Further discussion of the optimal combinations and interactions.
effect of specimen geometry related to standard development is pro­
vided in Section 5.1. 4.8. Other parameters

4.6. Specimen fabrication method Other, less commonly studied, parameters include flow multiplier
[107], printing temperature and speed [98,106,107], infill percentage
Three studies [46,47,81] investigated the effect of various fabrica­ [76,79,106], and infill pattern [76]. Tensile testing results of 100 %
tion methods. In [46], the UTS of polished specimens (prepared by infill specimens are affected significantly by the quality of interlayer and
grinding the surface) was higher than that of as-built specimens. The intralayer bonds. Thus, translating strength and stiffness values of a solid
specimen surface was polished to avoid crack initiation at the specimen specimen to that of a part with an entirely different raster structure
perimeter, thus enabling the authors to examine the effects of layer would lead to misguided and unreliable uses of these tensile properties.
height independent of surface stress concentrations. Lower layer heights Therefore, it is important that infill percentage and patterns reflect those
already result in reduced surface roughness and therefore polishing had of the final part.
a smaller effect on the UTS. There was minimal difference in modulus Four studies investigated the type of 3D printer [79,96,103,105], but
between polished and as-built specimens. On the contrary, water jet more research should investigate this parameter since understanding its
cutting eliminated perimeter contours, which resulted in significantly effects will be important for translating standardization from
lower UTS and modulus than as-built specimens in [47] and [81]. The industry-level to low-cost machines. 3D printers vary considerably in
remaining studies did not investigate the effect of fabrication methods: their ability to insulate the build chamber, which affects the cooling rate
two studies [78,84] fabricated all specimens by machining them from an of each printed layer [33], [47] and alters the polymer diffusion of
extruded sheet, while the remaining studies extruded the specimens adjacent rasters.
directly to size. The latter fabrication method is clearly the most com­
mon; however, test specimens prepared with this method may not yield 4.9. Interaction effects and combination optimization
results that accurately represent the final part, which is discussed further
in Section 5.1. Failing to address effects from interacting parameters may lead to
misinterpretation of results [48,98,106]. Using results from an ANOVA
4.7. Layer height and raster width analysis, Rankouhi et al. [94] performed a Tukey-Kramer test and found
that reducing layer height at a constant raster angle of 0˚had a much
Raster width and layer height affect the void density in printed parts, greater effect on improving UTS compared to reducing layer height at a
as wider rasters and lower layer heights reduce void size between rasters constant raster angle of 90˚. Lanzotti et al. [36] performed a robust
(Fig. 3). Of the studies listed in Table 3, 14 of 37 investigated layer surface regression analysis among layer height, raster angle, and num­
height effects. Three of these only looked at two different layer heights ber of contours. They found that, at 0˚raster angle, changing the number
[79,86,94], and all found that the lower layer height resulted in greater of contours had very little effect on UTS and modulus; however, as raster
UTS. The remaining 11 publications looked at three or more layer angle increased to 90˚, the number of contours had an increasingly
heights: seven found that the lowest layer height resulted in greater UTS greater effect.
[46,78,83,84,95,103,106]; three found that UTS increased to a peak, Testing more than two levels per variable is often impractical due to
then decreased as layer height increased [36,93,107]; and one found limited number of resources, effort, and time. Design of experiments
that UTS decreased to a minimum, then increased as layer height (DOE) methods can help identify optimal parameter combinations that
increased [109]. Justifications for lower layer heights resulting in may not otherwise be detected solely from experimental studies [42].
greater UTS were generally based on the reduced void density and Nonetheless, a rigorous testing of multiple levels per variable was

Fig. 3. Effect of raster width and layer height on void density. The right drawing has a larger raster width and lower layer height compared to the left drawing.

14
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

performed by Laureto and Pearce [107] who leveraged 47 different 5. Proposed guidelines for coupon preparation
printers and operators at their research facility to fabricate 423 speci­
mens spanning nine different layer heights, 12 printing speeds, 11 In this section, we focus on identifying the consistencies and in­
printing temperatures, and five flow rate multipliers. Unfortunately, consistencies in specimen preparation methods from the literature
interactive effects among these parameters were not investigated. identified in Table 3, and propose guidelines for implementing stan­
DOE methods were used in four of 37 publications [36,86,98,109] to dardized methods. These guidelines, presented in Sections 5.1 through
determine optimal parameter combinations without carrying out full 5.6, follow a similar chronological order to the specimen preparation
factorial experiments. The UTS was predicted using a Group Method for procedure (Fig. 4): preparing the CAD file, selecting the slicing param­
Data Handling in [86], and a central composite design in [98] and [109]. eters, then fabricating and inspecting the specimen.
While these models showed good correlation to experimental runs, the
optimal parameters found in the model for maximum UTS were not 5.1. Specimen fabrication procedure and geometry
explicitly validated experimentally. In [36], however, optimal parame­
ters for maximum UTS and modulus were predicted using a central Prior to selecting any process parameter, the test specimen fabrica­
composite design, and this combination was validated experimentally, tion procedure and geometry must be identified. In most previous
where the resulting UTS and modulus were 6 % and 10.9 % lower than studies, the coupon specimens were extruded to their exact size, which is
predicted, respectively. Experimental validation of numerical models is surprising since the preparation method provided in ASTM D638, ISO
important not only for the rigor of research, but to ensure the identified 527–2 and ASTM D3039 recommends to machine the specimen geom­
parameters are in fact achievable for the precision and specifications of a etry from a sheet of material. The justification may be that machining
given MEX printer and controllability of the slicing software. the perimeters results in a different surface finish than that of the final
part, as discussed in Section 3.2. A few exceptions were found. Garg and
Bhattacharya [84] extruded rectangular specimens using MEX, then
4.10. Implications for this review and standardization development trimmed them to their exact geometry. Allum et al. [78] developed a
toolpath strategy to control the extrusion of a four-sided hollow box with
The literature shows that mechanical properties such as tensile single-filament wide walls. The specimen geometry was then cut from
strength and stiffness are dependent on build orientation, geometry, the walls of the box in two orientations (YZ and Z) using a custom razor
layer height, raster angle, fabrication method, and other parameters. blade tool. The authors stated that this preparation procedure enabled
Understanding these trends is important and helps inform design pro­ better control of "nozzle position, extrusion volume, speed and sequence
cesses. However, due to differences in testing methodologies, some of deposition”, thus creating tensile specimens that more consistently
disagreements in the findings remain, and direct comparisons between failed in the gauge region. Zhang et al. [47] stated that using water jet
studies on mechanical properties of MEX specimens are not possible. cutting minimizes the amount of heat generated during the machining
Collectively the investigations presented in this paper imply that process, thus helping to preserve the mesostructure of the specimen
neglecting the effects of these parameter settings when preparing test perimeter. However, contrary to this conjecture, water jet cut specimens
specimens may result in inaccurate values for the mechanical properties demonstrated wider variability in results and were significantly weaker
of MEX printed parts, and hence premature failure for parts designed and less stiff than their as-built counterparts. The authors attributed this
based on these properties. It is, therefore, crucial to standardize to the lack of a contour, explaining that contours reduce variance in
parameter considerations and fabrication methods for realizing consis­ mechanical properties by homogenizing the strain distribution along the
tency. An in-depth discussion follows on how these parameters may be longitudinal axis. This effect on variance was also observed in [81].
considered in the development of a new AM standard. Moreover, Striemann et al. [46] demonstrated that polishing the

Fig. 4. Approach for organizing the proposed standard guidelines and recommendations.

15
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

specimen perimeter may yield more accurate interlayer material prop­ taken by Torrado and Roberson [77] who investigated whether XY
erties. These may be suitable specimen preparation procedures, pro­ specimens with zero contours and 90˚rasters could be used to simulate Z
vided that the machined or polished specimens appropriately reflect the specimens. They found that the two orientations may result in equiva­
final part. Nonetheless, only a limited number of studies have investi­ lent tensile results. However, more research is needed to validate this
gated the effects of fabrication methods on MEX mechanical properties, finding. Although both specimens consist of filaments laid down trans­
and a better understanding of these effects will help inform a standard verse to the direction of subsequent loading, the orientation of the
specimen preparation procedure. elliptical cross-section of the individual filaments is different.
Regardless of the fabrication procedure, the final geometry must be The main directions of loading relative to the raster angles and build
standardized. The studies reviewed in Table 3 used 16 different geom­ orientation of the final part should be identified and evaluated through
etries: ASTM Type I, Type II, Type IV, Type V, ISO 527–2, ISO 20753 testing specimens designed with raster angles and build orientations that
A12, ASTM D3039, ASTM D1708, DIN 53442, GB/T 16421, and six simulate the same critical loading conditions. If there is any indication
different modifications of these standard geometries (Fig. 5). A major that a part will undergo loading transverse to layer deposition, the
reason for investigating various geometries was to ensure specimen prospective AM standard must require testing of Z-oriented specimens to
failure in the acceptable gauge region, i.e., within the tapered region of a determine the important limiting Z-strength. The standard should also
dog-bone specimen, or at least one width-distance from the grip of a provide guidelines for utilizing support structures to facilitate controlled
rectangular specimen [36,75,79,97]. Croccolo et al. [75] and Lanzotti fabrication in the vertical direction, with careful attention given to avoid
et al. [36] specified modified geometries that resulted in acceptable the gauge region when designing the support.
failure locations, and used these geometries for their respective studies.
Zhang et al. [97] found acceptable failure locations with ASTM D638 5.3. Slicing process parameters
Type II and ASTM D3039 specimen types, and consistently unacceptable
failure with ASTM D638 Type I. They explained, similarly to other Contrary to conventional material testing (where the homogenous
studies [36,75,82], that failure outside the gauge region of a dog-bone specimen can represent the final part without any special preparation
specimen is likely due to the increased stress concentration around its considerations), it is critical that an MEX testing standard specifies that
filleted region, which is greatest with Type I. Zhang et al. [47] calculated the heterogenous test specimens and final part be fabricated using the
the stress concentration factor (using an empirically derived equation by same slicing parameters. These parameters include, but are not limited
[110]) for 12 geometries: ISO 527–2 Type 1A and 1B, ASTM D638 Types to, raster angle, raster width, layer height, air gap, infill percentage,
I, II, III IV and IV, and ISO 20753 Types A12, A13, A14, A15 and A18. infill pattern, and number of contours. It is relatively simple to translate
They concluded that ASTM D638 Type II geometry has the lowest stress these meso- and macroscale parameters from the test specimen to the
concentration factor. Furthermore, Miller et al. [79] found best consis­ final part; however, we address four notable points to facilitate this:
tency in gauge-region failure with rectangular ASTM D3039 specimens
compared to ASTM D638 Type I and Type IV dog-bones (they did not • Due to the anisotropy dependent on raster orientation in each layer,
test ASTM D638 Type II). These findings may suggest that MEX speci­ classical laminate theory (CLT) may be used to select optimal raster
mens should be rectangular (with added tabs to protect the grip region) orientations for the intended design and loading cases. CLT was
rather than the traditional dog-bone shape, to avoid added stress con­ validated by Choi and Kortschot [112] for the prediction of stiffness
centrations. However, as there is no consensus on the most appropriate and Poisson’s ratio for fiber-reinforced MEX parts, and they
specimen geometry, this topic requires further study. demonstrated that this model can be adjusted for the effect of con­
tours. CLT may also be applied to neat MEX specimens (i.e., no fiber
5.2. Orientation specifications reinforcement) due to the anisotropy caused from layer deposition,
despite the type of filament used. CLT was used in [91] and [100],
Despite its strong effect on material properties, build orientation was showing good results in estimating strength and stiffness, and Li et al.
tested in only 11 of the 37 publications identified in Table 3. Zaldivar [113] was one of the first to validate the CLT model while consid­
et al. [101] provided the most extensive investigation on this parameter ering void density of the MEX specimens. It is reasonable to foresee
by building specimens in six different orientations. Z-oriented specimens the use of CLT or other analytical models as supporting tools used in
clearly exhibit weaker tensile strengths and less ductility. Therefore, the slicing programs to generate optimal raster orientations for the
material properties in this limiting and unfavorable orientation should anticipated loading, rather than resorting to default parameters. It
be understood to account for worst-case loading of the final part. will be important for the user to understand the reasoning behind
Findings from [64,78] suggest that the poor performance of Z-oriented raster orientation selection, and ensure the test specimen accurately
specimens is due to the presence of surface notches, which result in an reflects the resulting anisotropy of the intended end-use part.
under-measured load-bearing area, and not due to weaker interlayer • The current practice for testing AM specimens is to use 100 % infill,
bonding, as the majority of literature suggests. This indicates that a with the exception of three studies [76,79,106]; however, this is not
method for accurately measuring the true cross-sectional area of the common practice for final parts. Procedures for testing specimens
specimen would be an important addition in an MEX standard. with less than 100 % infill may require special attention due to their
Fabricating specimens in the Z orientation may be difficult due to discontinuity, making current test methods like ASTM D638 and ISO
their high height-to-width aspect ratio. Casavola et al. [91] stated that 527 inapplicable as they apply to homogenous materials [50].
they did not test in the Z orientation due to the mechanical instability of Scalability of infill patterns between test specimens and final parts
the thin vertical specimen, and Dietrich and Hayes [111] have indicated may also be problematic if the same infill pattern is not used. With
that the AM industry lacks a procedure for 3D printing vertical coupons. infinite infill patterns available, more research is needed to investi­
The literature offers three main techniques for enabling fabrication in gate the effect of infill percentages and patterns.
the Z orientation: (1) adding surrounding support material [87,108], (2) • A negative air gap may be referred to as an infill overlap percentage
machining vertical coupons from a stable upright structure, such as a in some slicing software [114,115]. The ability to set positive air
hollow box [78], or (3) adding a sacrificial rib in the same material at the gaps may only be available in robust slicing software such as Insight
base grip region [77,107,111]. Once removed, sacrificial ribs or support by Stratasys [116], which allows the user to control the air gap be­
material may leave surface defects; however, the gauge region will tween contours and rasters, as well as between adjacent rasters. This
remain unaffected if they are located only in the grip region. Caution review does not discuss existing slicing software for MEX, but this
should be taken with these methods to avoid adding stress concentra­ should be an important consideration in standardization as it directly
tions and interfering with the gauge region. An interesting approach was controls how the part is fabricated [29], [117].

16
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Fig. 5. Sixteen different test specimen geometries used for tensile testing identified in this review. One geometry possibility is shown for ASTM D3039; however, note
that the standard allows for various other width and length dimensions if similar aspect ratios are achieved. Non-standard geometries are sourced from Garg and
Bhattacharya [84], Sebert et al. [96], Lanzotti et al. [36], Croccolo et al. [75], Onwubolu and Rayegani [86], and Nomani et al. [95].

17
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

• The selection of contour lines in the test specimens must also be location, and results were more representative of the bulk of the part.
representative of those found in the final part. Best judgements A standard test method should recommend strategic placement of
should be made, based on the function and loading of the final part, the parts on the build platform, sequential printing of test specimens, or
to avoid fractures that typically propagate from the outer surface (i. manually adjusting the toolpath algorithm to avoid having the print
e., from the contour filaments). Additionally, to reduce stress con­ head travel across other specimens. To reduce oozing, a faster retraction
centrations from notches on the surface due to transverse filament may be used. Seam control may also be enabled, if it is a controllable
deposition, at least one contour should be used. setting in the slicing software, to start and stop each layer at a specified
location on each part. For X and YZ coupons, it would be best to place
the seams within the grip region, particularly at the corners so that the
5.4. Toolpath optimization nozzle avoids cross-over. However, for Z coupons, one cannot avoid
having the seam in the gauge region. Fig. 7 demonstrates how failure can
Once slicing parameters are set, the next process is to generate the occur precisely at the seam location on specimens. Hossain et al. [108]
toolpaths for each layer. After the first inspection of the generated observed that their Z coupons exhibited a more visible seam line, thus
toolpath, further iterations of changing slicing parameters and rein­ creating a poorer surface finish, which may have led to greater standard
specting the toolpath may be required. Hossain et al. [108] looked at the deviations in their results [108]. Therefore, we recommend tuning the
detailed rendered toolpaths generated by slicing software to identify air toolpath to vary its start and end points on each layer to limit accu­
gaps that arise in the cross-section due to the ratio of combined raster mulation of oozing material in one location, or have the toolpath start
width, contour width, and air gap, to the total width of the specimen. and end within the center of each layer’s cross-section to eliminate
They found that using default slicing parameters for raster width, con­ filament notches on the part’s outer surface. Koch et al. [81] stated that
tour width and raster-to-raster air gap resulted in the largest air gaps in their toolpath was set to start at the lower left corner of the specimen at
the toolpath between adjacent rasters and contours. The slicing pa­ each layer height. Research articles do not commonly report these
rameters were, therefore, precisely modified to minimize these air gaps. toolpath details in their methods section, but it should become common
However, optical imaging of the printed parts revealed that air gaps practice. Moreover, to eliminate inconsistencies in the toolpath due to
were still visible between contours and rasters. Through iterative the various slicing parameters, the slicing software may be avoided
testing, they determined that applying a − 0.013 mm air gap parameter completely by developing GCode programs to generate standard tool­
in the slicer resulted in 3D printed specimens with no gaps. Since air gap paths, as utilized in [78]. In fact, different slicing software have been
has a significant effect on material properties [82,86,100,108], using a shown to output specimen toolpaths with varying mechanical proper­
technique of this kind to, at least, identify whether air gaps exist, would ties, even though input slicing parameters were the same [118]. While a
be valuable and important to disclose in the test results. The work by standard GCode framework for specimen preparation is intriguing, it
Hossain et al. [108] demonstrates the need to not only optimize the must be accompanied by a set of instructions to guide designers or en­
toolpath, but to also validate the toolpath by inspecting the printed parts gineers in modifying the code to create custom specimens that represent
in a separate calibration procedure. This will be discussed in Section 5.6 the final part structure. A standard GCode framework would eliminate
for inspection and validation. variabilities resulting from slicer programs, and certain toolpath
To improve experimental efficiency, multiple test specimens may be parameters.
3D printed simultaneously. Where possible, the toolpath should be
analyzed and controlled to reduce the amount of travel across adjacent
parts. Travel paths can be inspected and controlled in the slicing soft­ 5.5. Printer and material specifications
ware (Fig. 6A). When the nozzle moves across the previous layer from
one coupon to the next, it may ooze (or “string”) some material, thus Where possible, materials should be procured from certified sup­
depositing an unknown amount of additional material and affecting the pliers (e.g., suppliers who adhere to ISO/ASTM 52903), to increase the
uniformity of the specimens (Fig. 6C). This printing defect can affect the level of certainty in fabricated parts, as also suggested by Garcia-
failure location during coupon testing, demonstrated in Fig. 6B, where Dominguez et al. [50]. Printing and build plate temperatures are most
failure occurs precisely in the same location as the travel path for often recommended by the material manufacturer, and should be
specimens 2–5. For specimen 1, however, the travel path did not cross adhered to. Incomplete polymer fusion may occur if the selected tem­
the narrow region of the specimen, and thus did not influence failure perature is too low, leading to void formation, while temperatures that

Fig. 6. Effect of toolpath on failure location. (A) Toolpath generated in the slicing software (ideaMaker). Dark blue lines indicate the travel path across each test
specimen. Red arrows indicate the specific travel paths that corresponded to the failure location of the test specimen. (B) Test specimens after failure, and (C) image
taken during 3D printing. Blue arrow shows the oozing material along the travel path for specimen 2.

18
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

Fig. 7. Effect of seam on failure location. (A) YZ specimens in the slicing software (GrabCAD) displaying the seam line. (B) Specimens after tensile testing,
demonstrating failure at the seam location.

are too high may lead to thermal aging or filament degradation [119]. quality and increasing the certainty of tensile test results. This involves a
The AM standard should require that the same material and printer visual inspection for sub-optimal layers, then weighing the sample for
intended for final part fabrication be used for the test specimens. under-extrusion. Nomani et al. [95] employed a similar method to
The thermal history present in MEX specimens from heating and determine void density, which was used to inform the interpretation of
cooling cycles of the layer-by-layer material deposition has a significant results from tensile and compressive tests. Riddick et al. [88] spoke to
impact on part characteristics [33,54]. A complex thermal profile forms, the quality of coupons and the importance of reducing void density to
where interfacial temperature periodically reaches its peak value from improve desired structural properties. To avoid volumetric error, they
the nozzle, and then decreases to the temperature of the build plate (for employ a strategy proposed by Masood et al. [121] to fill in the sliced
layers close to the build plate) or the temperature of the build chamber cross sections with the most material possible, therefore minimizing
(for layers farther from the build plate). Only while the interfacial gaps. Cwikla et al. [76] found poorer surface quality when using a fill
temperature is above the glass transition temperature can polymer multiplier outside of 0.9–1.1, and Onwubolu et al. [86] found that an air
diffusion occur. Therefore, reducing the cooling time between layers gap of − 0.001 mm resulted in an “over-fill” of the material, and nega­
will increase bonding strength [33], [54]. The time delay between layers tively impacted the surface quality.
was purposely controlled in [78] to ensure that all specimens were Validating the quality of test specimens is also contingent upon a
exposed to similar thermal variability throughout the part. Ultimately, minimum number of samples meeting the desired criteria prior to me­
the time scale between layers in the final part should be estimated and chanical testing. The number of samples tested in the identified publi­
reflected in the preparation of the test specimens. Since final parts likely cations varied from 2 to 35, with five samples being the most common,
have larger cross-sectional areas than their corresponding test specimen, as recommended by ASTM D638 and ISO 527. Only 16 of 37 studies used
a delay would be needed when fabricating specimens. This can be statistical analysis methods to report significant effects from various
achieved by implementing a delay in the GCode, or by printing a dummy parameters on strength and modulus values. Furthermore, not only
tower next to the specimens to increase the time spent printing on each should enough specimens be inspected for quality prior to testing, but a
layer, and thus achieve more cooling between layers. statistically relevant number of samples must fail in the accepted gauge
As discussed by Dey and Yodo [42], realization of a standardized region for results to be valid. Miller et al. [79] suggested that more
MEX process may be facilitated by integrating closed-loop feedback samples than currently required by ASTM and ISO standards may be
systems and machine learning algorithms into the MEX equipment to needed due to inconsistent failure locations in MEX specimens [79], and
monitor the process performance, such as the detection of temperature Allum et al. [78] proposed a novel preparation procedure that may
fluctuations or defect formations. Deviations from process parameters enable consistency in failure location.
may be flagged and corrected with an advanced control system. This
may be an important area of future research to guide standardization 6. Conclusions
development.
A standardized testing protocol does not yet exist for MEX specimens,
and the methods used by researchers and manufacturers at present
5.6. Inspection and validation of coupon preparation
therefore vary widely and cannot be expected to produce consistent
results. Standardization is a challenging task, especially for a continually
As demonstrated by Hossain et al. [108], discrepancies exist between
changing field such as AM. This paper identified the extent of disparity
rendered toolpaths and actual toolpaths in the fabricated part. These
in MEX tensile test specimen preparation methods used in literature and
discrepancies should be reduced, or at least identified and quantified
the resulting effects on material properties such as ultimate tensile
where possible. Tymrak et al. [103] pointed out that for low-cost,
strength and modulus. We have highlighted opportunities and proposed
open-source 3D printers, exact air gaps between rasters cannot be
guidelines for creating a standard method for test specimen preparation.
specified prior to printing, and the actual positive or negative air gaps
These guidelines relate to the dimensions and fabrication of test speci­
may vary between printers depending on printer technology. Thus, it
mens, build and raster orientation, slicing parameters, toolpath opti­
may be necessary to perform validation through experimental
mization, printer and material specifications, and inspection of
measurements.
fabricated specimens. Consideration of these specifications in a stan­
Various authors have described methods of inspection for void
dardized method may facilitate comparisons between published results
density and overall quality of test specimens. Tanikella et al. [120]
and support progression of MEX technology.
provided a two-step process for solving the challenge of poor print

19
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

CRediT authorship contribution statement [21] ASTM International, Additive manufacturing - General principles - Fundamentals
and vocabulary. ISO/ASTM 52900, 2021.
[22] O. Abdulhameed, A. Al-Ahmari, W. Ameen, S.H. Mian, Additive manufacturing:
Azhari Fae: Writing – review & editing, Conceptualization. Phillips challenges, trends, and applications, Adv. Mech. Eng. 11 (2) (2019) 1–27, https://
Clara: Writing – review & editing, Writing – original draft, Data cura­ doi.org/10.1177/1687814018822880.
tion, Conceptualization. Kortschot Mark: Writing – review & editing. [23] B.T. Wittbrodt, et al., Life-cycle economic analysis of distributed manufacturing
with open-source 3-D printers, Mechatronics 23 (6) (2013) 713–726, https://doi.
org/10.1016/J.MECHATRONICS.2013.06.002.
Declaration of Competing Interest [24] G.A. Nowlan, Developing and implementing 3D printing services in an academic
library, Libr. Hi Tech. 33 (4) (2015) 472–479, https://doi.org/10.1108/LHT-05-
2015-0049.
The authors declare that they have no known competing financial [25] S. Terekhina, T. Tarasova, S. Egorov, L. Guillaumat, M.L. Hattali, On the
interests or personal relationships that could have appeared to influence difference in material structure and fatigue properties of polyamide specimens
produced by fused filament fabrication and selective laser sintering, Int. J. Adv.
the work reported in this paper. Manuf. Technol. 111 (1–2) (2020) 93–107, https://doi.org/10.1007/S00170-
020-06026-X.
Data availability [26] J.H. Tan, W.L.E. Wong, K.W. Dalgarno, An overview of powder granulometry on
feedstock and part performance in the selective laser melting process, Addit.
Manuf. 18 (2017) 228–255, https://doi.org/10.1016/J.ADDMA.2017.10.011.
No data was used for the research described in the article. [27] C.A. Chatham, T.E. Long, C.B. Williams, A review of the process physics and
material screening methods for polymer powder bed fusion additive
manufacturing, Prog. Polym. Sci. 93 (2019) 68–95, https://doi.org/10.1016/J.
Acknowledgements PROGPOLYMSCI.2019.03.003.
[28] W.M. Johnson, M. Rowell, B. Deason, M. Eubanks, Comparative evaluation of an
University of Toronto – Dean’s Spark Professorship Canada, Grad­ open-source FDM system, Rapid Prototyp. J. 20 (3) (2014) 205–214, https://doi.
org/10.1108/RPJ-06-2012-0058.
uate Scholarships – Master’s program. [29] S. Wickramasinghe, T. Do, P. Tran, FDM-based 3D printing of polymer and
associated composite: a review on mechanical properties, defects and treatments,
References Polymers) 12 (7) (2020) 1529, https://doi.org/10.3390/POLYM12071529.
[30] C.C. Spackman, C.R. Frank, K.C. Picha, J. Samuel, 3D printing of fiber-reinforced
soft composites: process study and material characterization, J. Manuf. Process.
[1] A.L. Russell, Standardization in history: a review essay with an eye to the future,
23 (2016) 296–305, https://doi.org/10.1016/J.JMAPRO.2016.04.006.
in: S. Bolin (Ed.), The Standards Edge: Future Generations, Sheridan Press, Ann
[31] N. van de Werken, H. Tekinalp, P. Khanbolouki, S. Ozcan, A. Williams,
Arbor, MI, 2005, pp. 247–260.
M. Tehrani, Additively manufactured carbon fiber-reinforced composites: state of
[2] M. Boulanger, M.E. Johnson, S.N. Luko, Reviews of Standards and Related
the art and perspective, Addit. Manuf. 31 (2020), 100962, https://doi.org/
Material: Statistical Standards and ISO, Part 1, Qual. Eng. 24 (1) (2012) 94–101,
10.1016/J.ADDMA.2019.100962.
https://doi.org/10.1080/08982112.2012.623956.
[32] X. Wang, M. Jiang, Z. Zhou, J. Gou, D. Hui, 3D printing of polymer matrix
[3] W. Blum, History of ASTM Committee B-8. In: Proceedings of the Symposium on
composites: a review and prospective, Compos. Part B Eng. 110 (2017) 442–458,
Properties, Tests, and Performance of Electrodeposited Metallic Coatings, 1957,
https://doi.org/10.1016/J.COMPOSITESB.2016.11.034.
doi:10.1520/STP46851S.
[33] X. Gao, S. Qi, X. Kuang, Y. Su, J. Li, D. Wang, Fused filament fabrication of
[4] M. Boulanger, M. Johnson, C. Perruchet, P. Thyregod, M. Boulanger, Evolution of
polymer materials: a review of interlayer bond,”, Addit. Manuf. 37 (2021),
international statistical standards via life cycle of products and services, Int. Stat.
101658 https://doi.org/10.1016/J.ADDMA.2020.101658.
Rev. 67 (2) (1999) 151, https://doi.org/10.2307/1403396.
[34] M. Seifi, et al., Progress towards metal additive manufacturing standardization to
[5] K.N. Mathes, ASTM Committee D09 - A History of Success. In: Proceedings of the
support qualification and certification, JOM 69 (3) (&;;2017) 439–455, https://
Symposium of Electrical Insulating Materials: International Issues, 2000, no.
doi.org/10.1007/S11837-017-2265-2.
1376, pp. 3–12, doi:10.1520/STP13449S.
[35] M. Gorelik, Additive manufacturing in the context of structural integrity, Int. J.
[6] ASTM, Standard Test Method for Tensile Properties of Plastics. ASTM D638-84,
Fatigue 94 (2017) 168–177, https://doi.org/10.1016/J.IJFATIGUE.2016.07.005.
ASTM, 1984.
[36] A. Lanzotti, M. Grasso, G. Staiano, M. Martorelli, The impact of process
[7] ASTM, Tensile Properties of Composite, Fiber Resin. ASTM D3039-76R89, ASTM,
parameters on mechanical properties of parts fabricated in PLA with an open-
1989.
source 3-D printer, Rapid Prototyp. J. 21 (5) (2015) 604–617, https://doi.org/
[8] J. Moylan, ASTM D3039 Considerations for Composite Tensile Testing. Element,
10.1108/RPJ-09-2014-0135.
Oct. 22, 2021. 〈https://www.element.com/nucleus/2017/astm-d3039-testin
[37] R.A. García-León, J.A. Gómez-Camperos, H.Y. Jaramillo, Scientometric review of
g-considerations〉 (accessed Jan. 30, 2022).
trends on the mechanical properties of additive manufacturing and 3D printing,
[9] M.H. Sherif, When is standardization slow? in: K. Jakobs (Ed.), Advanced Topics
J. Mater. Eng. Perform. 30 (7) (2021) 4724–4734, https://doi.org/10.1007/
in Information Technology Standards and Standardization Research, vol. 1 IGI
S11665-021-05524-7.
Global, 2006, pp. 128–137.
[38] E. Cuan-Urquizo, E. Barocio, V. Tejada-Ortigoza, R.B. Pipes, C.A. Rodriguez,
[10] E.A. Campo, Polymeric Materials and Properties, in: E.A. Campo (Ed.), Selection
A. Roman-Flores, Characterization of the mechanical properties of FFF structures
of Polymeric Materials, William Andrew Publishing, 2008, pp. 1–39.
and materials: a review on the experimental, computational and theoretical
[11] M. Attaran, The rise of 3-D printing: the advantages of additive manufacturing
approaches, Materials 12 (6) (2019), https://doi.org/10.3390/MA12060895.
over traditional manufacturing, Bus. Horiz. vol. 60 (5) (. 2017) 677–688, https://
[39] J.R.C. Dizon, A.H. Espera, Q. Chen, R.C. Advincula, Mechanical characterization
doi.org/10.1016/J.BUSHOR.2017.05.011.
of 3D-printed polymers, Addit. Manuf. 20 (2018) 44–67, https://doi.org/
[12] B. Stucker, Additive Manufacturing Technologies: Technology Introduction and
10.1016/J.ADDMA.2017.12.002.
Business Implications. Front. Eng. Reports Leading-Edge Eng. From 2011 Symp.
[40] D. Popescu, A. Zapciu, C. Amza, F. Baciu, R. Marinescu, FDM process parameters
Natl. Acad. Press. Washingt. DC, pp. 5–14, Sep. 2011, doi:10.17226/13274.
influence over the mechanical properties of polymer specimens: a review, Polym.
[13] J.C. Najmon, S. Raeisi, A. Tovar, Review of additive manufacturing technologies
Test. 69 (2018) 157–166, https://doi.org/10.1016/J.
and applications in the aerospace industry, in: F. Froes, R. Boyer (Eds.), Additive
POLYMERTESTING.2018.05.020.
Manufacturing for the Aerospace Industry, Elsevier, 2019, pp. 7–31.
[41] M. Ahmadifar, K. Benfriha, M. Shirinbayan, A. Tcharkhtchi, Additive
[14] J.C. Vasco, Additive manufacturing for the automotive industry, in: J. Pou,
manufacturing of polymer-based composites using fused filament fabrication
A. Riveiro, P. Davim (Eds.), Additive Manufacturing, Elsevier, 2021, pp. 505–530.
(FFF): a review, Appl. Compos. Mater. 28 (2021) 1335–1380, https://doi.org/
[15] D. Beiderbeck, H. Krüger, T. Minshall, The future of additive manufacturing in
10.1007/S10443-021-09933-8.
sports, in: S.L. Schmidt (Ed.), Future of Business and Finance, 21st Century Sports,
[42] A. Dey, N. Yodo, A systematic survey of FDM process parameter optimization and
Springer, Cham, 2020, pp. 111–132.
their influence on part characteristics, J. Manuf. Mater. Process. 3 (3) (2019) 64,
[16] R. Kumar, M. Kumar, J.S. Chohan, The role of additive manufacturing for
https://doi.org/10.3390/JMMP3030064.
biomedical applications: a critical review, J. Manuf. Process. 64 (2021) 828–850,
[43] ASTM International, Standard Test Method for Tensile Properties of Plastics.
https://doi.org/10.1016/J.JMAPRO.2021.02.022.
ASTM D638–14, 2014.
[17] T.M. Morrison, P. Pathmanathan, M. Adwan, E. Margerrison, Advancing
[44] ASTM International, Plastics - Determination of tensile properties - Part 2: Test
regulatory science with computational modeling for medical devices at the FDA’s
conditions for moulding and extrusion plastics. ISO 527–2, 2012.
office of science and engineering laboratories, Front. Med. (2018) 241, https://
[45] R. Torre, S. Brischetto, Experimental characterization and finite element
doi.org/10.3389/FMED.2018.00241.
validation of orthotropic 3D-printed polymeric parts, Int. J. Mech. Sci. 219
[18] U.P. Breuer, Testing. Commercial Aircraft Composite Technology, Springer, 2016,
(2022), https://doi.org/10.1016/J.IJMECSCI.2022.107095.
pp. 133–140.
[46] P. Striemann, D. Huelsbusch, M. Niedermeier, F. Walther, Application-oriented
[19] A.N. Dickson, H.M. Abourayana, D.P. Dowling, 3D printing of fibre-reinforced
assessment of the interlayer tensile strength of additively manufactured polymers,
thermoplastic composites using fused filament fabrication – a review, Polymers)
Addit. Manuf. 46 (2021), 102095, https://doi.org/10.1016/J.
12 (10) (2020) 2188, https://doi.org/10.3390/POLYM12102188.
ADDMA.2021.102095.
[20] J. Wang, H. Xie, Z. Weng, T. Senthil, L. Wu, A novel approach to improve
mechanical properties of parts fabricated by fused deposition modeling, Mater.
Des. 105 (2016) 152–159, https://doi.org/10.1016/J.MATDES.2016.05.078.

20
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

[47] Y. Zhang, J.P. Choi, S.K. Moon, Effect of geometry on the mechanical response of [74] M. Jin, et al., Filament materials screening for FDM 3D printing by means of
additively manufactured polymer, Polym. Test. 100 (2021), 107245, https://doi. injection-molded short rods, Macromol. Mater. Eng. 303 (12) (2018), 1800507,
org/10.1016/J.POLYMERTESTING.2021.107245. https://doi.org/10.1002/MAME.201800507.
[48] A.M. Forster, Materials testing standards for additive manufacturing of polymer [75] D. Croccolo, M. De Agostinis, G. Olmi, Experimental characterization and
materials: state of the art and standards applicability, NIST Interagency/Internal analytical modelling of the mechanical behaviour of fused deposition processed
Report (NISTIR), National Institute of Standards and Technology, Gaithersburg, parts made of ABS-M30, Comput. Mater. Sci. 79 (2013) 506–518, https://doi.org/
MD, [online], 2015, https://doi.org/10.6028/NIST.IR.8059. 10.1016/J.COMMATSCI.2013.06.041.
[49] A. Garcia-Dominguez, J. Claver, A.M. Camacho, M.A. Sebastian, Considerations [76] G. Ćwikła, C. Grabowik, K. Kalinowski, I. Paprocka, P. Ociepka, The influence of
on the applicability of test methods for mechanical characterization of materials printing parameters on selected mechanical properties of FDM/FFF 3D-printed
manufactured by FDM, Materials 13 (1) (2019) 28, https://doi.org/10.3390/ parts, IOP Conf. Ser. Mater. Sci. Eng. 227 (1) (2017), 012033, https://doi.org/
MA13010028. 10.1088/1757-899X/227/1/012033.
[50] A. Garcia-Dominguez, J. Claver, A.M. Camacho, M.A. Sebastian, Analysis of [77] A.R. Torrado, D.A. Roberson, Failure analysis and anisotropy evaluation of 3D-
general and specific standardization developments in additive manufacturing printed tensile test specimens of different geometries and print raster patterns,
from a materials and technological approach, IEEE Access 8 (2020) J. Fail. Anal. Prev. 16 (1) (2016) 154–164, https://doi.org/10.1007/S11668-016-
125056–125075, https://doi.org/10.1109/ACCESS.2020.3005021. 0067-4.
[51] A. Kazemian, X. Yuan, E. Cochran, B. Khoshnevis, Cementitious materials for [78] J. Allum, A. Moetazedian, A. Gleadall, V.V. Silberschmidt, Interlayer bonding has
construction-scale 3D printing: laboratory testing of fresh printing mixture, bulk-material strength in extrusion additive manufacturing: new understanding
Constr. Build. Mater. 145 (2017) 639–647, https://doi.org/10.1016/J. of anisotropy, Addit. Manuf. 34 (2020), 101297, https://doi.org/10.1016/J.
CONBUILDMAT.2017.04.015. ADDMA.2020.101297.
[52] T. Letcher, B. Rankouhi, and S. Javadpour, Experimental study of mechanical [79] A. Miller, C. Brown, G. Warner, Guidance on the use of existing ASTM polymer
properties of additively manufactured abs plastic as a function of layer testing standards for ABS parts fabricated using FFF, Smart Sustain. Manuf. Syst. 3
parameters. In: Proceedings of the ASME’s International Mechanical Engineering (1) (2019) 122–138, https://doi.org/10.1520/SSMS20190051.
Congress & Exposition, vol. 2A-2015, 2015, doi:10.1115/IMECE2015-52634. [80] Wells D. Standardization in Additive Manufacturing: Challenges in Structural
[53] S. Guessasma, S. Belhabib, H. Nouri, S. Belhabib, Understanding the Integrity Assurance. Presented at: Symposium on Fatigue and Fracture of Additive
microstructural role of bio-sourced 3D printed structures on the tensile Manufactured Materials and Components, 2017.
performance, Polym. Test. 77 (2019), 105924, https://doi.org/10.1016/j. [81] C. Koch, L. Van Hulle, N. Rudolph, Investigation of mechanical anisotropy of the
polymertesting.2019.105924ï. fused filament fabrication process via customized tool path generation, Addit.
[54] T.J. Coogan, D.O. Kazmer, Bond and part strength in fused deposition modeling, Manuf. 16 (2017) 138–145, https://doi.org/10.1016/J.ADDMA.2017.06.003.
Rapid Prototyp. J. 23 (2) (2017) 414–422, https://doi.org/10.1108/RPJ-03- [82] S.H. Ahn, M. Montero, D. Odell, S. Roundy, P.K. Wright, Anisotropic material
2016-0050. properties of fused deposition modeling ABS, Rapid Prototyp. J. 8 (4) (2002)
[55] S.J. Park, et al., Tensile test of additively manufactured specimens with external 248–257, https://doi.org/10.1108/13552540210441166.
notch removed via laser cutting in material extrusion, Polym. Test. 110 (2022), [83] S.R. Rajpurohit, H.K. Dave, Effect of process parameters on tensile strength of
107581, https://doi.org/10.1016/J.POLYMERTESTING.2022.107581. FDM printed PLA part, Rapid Prototyp. J. 24 (8) (2018) 1317–1324, https://doi.
[56] Y. Song, Y. Li, W. Song, K. Yee, K.Y. Lee, V.L. Tagarielli, Measurements of the org/10.1108/RPJ-06-2017-0134.
mechanical response of unidirectional 3D-printed PLA, Mater. Des. 123 (2017) [84] A. Garg, A. Bhattacharya, An insight to the failure of FDM parts under tensile
154–164, https://doi.org/10.1016/J.MATDES.2017.03.051. loading: finite element analysis and experimental study, Int. J. Mech. Sci. 120
[57] P. Striemann, D. Huelsbusch, S. Mrzljak, M. Niedermeier, F. Walther, Systematic (2017) 225–236, https://doi.org/10.1016/J.IJMECSCI.2016.11.032.
approach for the characterization of additive manufactured and injection molded [85] C. Ziemian, M. Sharma, S. Ziemi, Anisotropic mechanical properties of ABS parts
short carbon fiber-reinforced polymers under tensile loading, Mater. Test. 62 (6) fabricated by fused deposition modelling, Mech. Eng. (2012), https://doi.org/
(2020) 561–567, https://doi.org/10.1515/MT-2020-620603. 10.5772/34233.
[58] S. Paul, Finite element analysis in fused deposition modeling research: A [86] G.C. Onwubolu, F. Rayegani, Characterization and optimization of mechanical
literature review, Measurement 178 (2021), 109320, https://doi.org/10.1016/J. properties of ABS parts manufactured by the fused deposition modelling process,
MEASUREMENT.2021.109320. Int. J. Manuf. Eng. 2014 (2014) 1–13, https://doi.org/10.1155/2014/598531.
[59] I. Gibson, D. Rosen, B. Stucker, M. Khorasani, Design for additive manufacturing. [87] I. Durgun, R. Ertan, Experimental investigation of FDM process for improvement
Additive Manufacturing Technologies, Springer, Cham, 2021, pp. 555–607. of mechanical properties and production cost, Rapid Prototyp. J. 20 (3) (2014)
[60] A. Bandyopadhyay, K.D. Traxel, Invited review article: Metal-additive 228–235, https://doi.org/10.1108/RPJ-10-2012-0091.
manufacturing – modeling strategies for application-optimized designs, Addit. [88] J.C. Riddick, M.A. Haile, R. Von Wahlde, D.P. Cole, O. Bamiduro, T.E. Johnson,
Manuf. 22 (2018) 758–774, https://doi.org/10.1016/J.ADDMA.2018.06.024. Fractographic analysis of tensile failure of acrylonitrile-butadiene-styrene
[61] A. du Plessis, et al., Beautiful and functional: a review of biomimetic design in fabricated by fused deposition modeling, Addit. Manuf. 11 (2016) 49–59, https://
additive manufacturing, Addit. Manuf. 27 (2019) 408–427, https://doi.org/ doi.org/10.1016/J.ADDMA.2016.03.007.
10.1016/J.ADDMA.2019.03.033. [89] N. Hill, M. Haghi, Deposition direction-dependent failure criteria for fused
[62] J. Plocher, A. Panesar, Review on design and structural optimisation in additive deposition modeling polycarbonate, Rapid Prototyp. J. 20 (3) (2014) 221–227,
manufacturing: towards next-generation lightweight structures, Mater. Des. 183 https://doi.org/10.1108/RPJ-04-2013-0039.
(2019), 108164, https://doi.org/10.1016/J.MATDES.2019.108164. [90] O.S. Es-Said, J. Foyos, R. Noorani, M. Mendelson, R. Marloth, B.A. Pregger, Effect
[63] S. Bhandari, R. Lopez-Anido, Finite element analysis of thermoplastic polymer of layer orientation on mechanical properties of rapid prototyped samples, Mater.
extrusion 3D printed material for mechanical property prediction, Addit. Manuf. Manuf. Process. 15 (1) (2000) 107–122, https://doi.org/10.1080/
22 (2018) 187–196, https://doi.org/10.1016/J.ADDMA.2018.05.009. 10426910008912976.
[64] J. Allum, A. Moetazedian, A. Gleadall, V.V. Silberschmidt, Discussion on the [91] C. Casavola, A. Cazzato, V. Moramarco, C. Pappalettere, Orthotropic mechanical
microscale geometry as the dominant factor for strength anisotropy in material properties of fused deposition modelling parts described by classical laminate
extrusion additive manufacturing, Addit. Manuf. 48 (2021), 102390, https://doi. theory, Mater. Des. 90 (2016) 453–458, https://doi.org/10.1016/J.
org/10.1016/J.ADDMA.2021.102390. MATDES.2015.11.009.
[65] ASTM International, Standard Test Method for Tensile Properties of Polymer [92] S. Ziemian, M. Okwara, C.W. Ziemian, Tensile and fatigue behavior of layered
Matrix Composite Materials. ASTM D3039, 2017. acrylonitrile butadiene styrene, Rapid Prototyp. J. 21 (3) (2015) 270–278,
[66] M. Friday, “A Comparison of Tension Test Data Using ASTM D 638 and ISO 527,” https://doi.org/10.1108/RPJ-09-2013-0086.
ASTM Spec. Tech. Publ., no. 1369, pp. 35–42, 2000, doi:10.1520/STP14340S. [93] W. Wu, P. Geng, G. Li, D. Zhao, H. Zhang, J. Zhao, Influence of layer thickness
[67] International Organization for Standardization, “ISO/TC 261 Business Plan,” Jul. and raster angle on the mechanical properties of 3D-printed PEEK and a
2020. Accessed: Aug. 31, 2021. [Online]. Available: 〈https://isotc.iso.org/li comparative mechanical study between PEEK and ABS, Materials 8 (9) (2015)
velink/livelink/fetch/2000/2122/687806/ISO_TC_261__Additive_manufacturin 5834–5846, https://doi.org/10.3390/MA8095271.
g_.pdf?nodeid=14655650&vernum=-2〉. [94] B. Rankouhi, S. Javadpour, F. Delfanian, T. Letcher, Failure analysis and
[68] ASTM International, Additive manufacturing - General principles - Part 3 Main mechanical characterization of 3D printed ABS with respect to layer thickness
characteristics and corresponding test methods. ISO 17296–3, 2014. and orientation, J. Fail. Anal. Prev. 16 (3) (2016) 467–481, https://doi.org/
[69] ASTM International, Standard Terminology for Additive 10.1007/S11668-016-0113-2.
Manufacturing—Coordinate Systems and Test Methodologies. ISO/ASTM 52921, [95] J. Nomani, D. Wilson, M. Paulino, M.I. Mohammed, Effect of layer thickness and
2019. cross-section geometry on the tensile and compression properties of 3D printed
[70] Z. Chen, L.-S. Turng, A review of current developments in process and quality ABS, Mater. Today Commun. 22 (2020), 100626, https://doi.org/10.1016/J.
control for injection molding, Adv. Polym. Technol. 24 (3) (2005) 165–182, MTCOMM.2019.100626.
https://doi.org/10.1002/ADV.20046. [96] A. Sebert, G. Bertacco, J. Nelson, Anisotropic mechanical performance of 3D
[71] W.E. Frazier, Metal additive manufacturing: a review, J. Mater. Eng. Perform. 23 printed polymers, TechConnect Briefs 4 (2017) 170–173.
(6) (2014) 1917–1928, https://doi.org/10.1007/S11665-014-0958-Z. [97] Y. Zhang, Y.C. Yeoh, G. Zheng, and S.K. Moon, Characterization of mechanical
[72] National Research Council, 3D printing in space, National Academies Press, properties of Ultem® 9085 using FDM. In: Proceedings of the 3rd International
Washington, DC, 2014, https://doi.org/10.17226/18871. Conference on Progress in Additive Manufacturing (Pro-AM 2018), (2018), pp.
[73] E. Sacco, S.K. Moon, Additive manufacturing for space: status and promises, Int. 451–457, doi:10.25341/D4D01W.
J. Adv. Manuf. Technol. 105 (10) (2019) 4123–4146, https://doi.org/10.1007/ [98] M. Ouhsti, B. El Haddadi, S. Belhouideg, Effect of printing parameters on the
S00170-019-03786-Z. mechanical properties of parts fabricated with open-source 3D printers in PLA by
fused deposition modeling, Mech. Mech. Eng. 22 (4) (2018) 895–907, https://doi.
org/10.2478/MME-2018-0070.

21
C. Phillips et al. Additive Manufacturing 58 (2022) 103050

[99] Ö. Keleş, C.W. Blevins, K.J. Bowman, Effect of build orientation on the build parameter modifications, J. Manuf. Sci. Eng. 136 (6) (2014), https://doi.
mechanical reliability of 3D printed ABS, Rapid Prototyp. J. 23 (2) (2017) org/10.1115/1.4028538.
320–328, https://doi.org/10.1108/RPJ-09-2015-0122. [109] A.K. Sood, R.K. Ohdar, S.S. Mahapatra, Parametric appraisal of mechanical
[100] J.F. Rodríguez, J.P. Thomas, J.E. Renaud, Mechanical behavior of acrylonitrile property of fused deposition modelling processed parts, Mater. Des. 31 (1) (2010)
butadiene styrene (ABS) fused deposition materials. Experimental investigation, 287–295, https://doi.org/10.1016/J.MATDES.2009.06.016.
Rapid Prototyp. J. 7 (3) (2001) 148–158, https://doi.org/10.1108/ [110] M.G. Garrell, A.J. Shih, E. Lara-Curzio, R.O. Scattergood, Finite-element analysis
13552540110395547. of stress concentration in ASTM D 638 tension specimens, J. Test. Eval. 31 (1)
[101] R.J. Zaldivar, D.B. Witkin, T. McLouth, D.N. Patel, K. Schmitt, J.P. Nokes, (2003) 52–57, https://doi.org/10.1520/JTE12359J.
Influence of processing and orientation print effects on the mechanical and [111] D.M. Dietrich and M.W. Hayes, “Z-axis test coupon structure and method for
thermal behavior of 3D-Printed ULTEM® 9085 material, Addit. Manuf. 13 (2017) additive manufacturing process,” US9109979B2, May 03, 2010.
71–80, https://doi.org/10.1016/J.ADDMA.2016.11.007. [112] J.Y. Choi, M.T. Kortschot, Stiffness prediction of 3D printed fiber-reinforced
[102] B. Wittbrodt, J.M. Pearce, The effects of PLA color on material properties of 3-D thermoplastic composites, Rapid Prototyp. J. 26 (3) (2020) 549–555, https://doi.
printed components, Addit. Manuf. 8 (2015) 110–116, https://doi.org/10.1016/ org/10.1108/RPJ-11-2018-0283/FULL/XML.
J.ADDMA.2015.09.006. [113] L. Li, Q. Sun, C. Bellehumeur, P. Gu, Composite modeling and analysis for
[103] B.M. Tymrak, M. Kreiger, J.M. Pearce, Mechanical properties of components fabrication of FDM prototypes with locally controlled properties, J. Manuf.
fabricated with open-source 3-D printers under realistic environmental Process. 4 (2) (2002) 129–141, https://doi.org/10.1016/S1526-6125(02)70139-
conditions, Mater. Des. 58 (2014) 242–246, https://doi.org/10.1016/J. 4.
MATDES.2014.02.038. [114] D. Brahm, “Cura.” Ultimaker, ver. 4.11.0, 2021.
[104] M.S. Uddin, M.F.R. Sidek, M.A. Faizal, R. Ghomashchi, A. Pramanik, Evaluating [115] Raise3D, “ideaMaker.” ver. 4.1.1.
mechanical properties and failure mechanisms of fused deposition modeling [116] Stratasys™ Support Center, “Insight Software Downloads.” 〈https://support.strat
acrylonitrile butadiene styrene parts, J. Manuf. Sci. Eng. 139 (8) (2017), https:// asys.com/en/resources/software-download〉 (accessed Aug. 28, 2021).
doi.org/10.1115/1.4036713. [117] I. Gibson, D. Rosen, B. Stucker, Additive Manufacturing Technologies: 3D
[105] J. Cantrell et al., Experimental characterization of the mechanical properties of Printing, Rapid Prototyping, and Direct Digital Manufacturing, second ed.,
3D printed ABS and polycarbonate parts. In: Proceedings of the Conference on Springer,, New York, 2015.
Society for Experimental Mechanics Series, 2017, vol. 3, pp. 89–105, doi: [118] E. Gkartzou, E.P. Koumoulos, C.A. Charitidis, Production and 3D printing
10.1007/978–3-319–41600-7_11. processing of bio-based thermoplastic filament, Manuf. Rev. 4 (2017) 1, https://
[106] X. Deng, Z. Zeng, B. Peng, S. Yan, W. Ke, Mechanical properties optimization of doi.org/10.1051/MFREVIEW/2016020.
poly-ether-ether-ketone via fused deposition modeling, Materials 11 (2) (2018), [119] C.G. Schirmeister, T. Hees, E.H. Licht, R. Mülhaupt, 3D printing of high density
https://doi.org/10.3390/MA11020216. polyethylene by fused filament fabrication, Addit. Manuf. 28 (2019) 152–159,
[107] J.J. Laureto, J.M. Pearce, Anisotropic mechanical property variance between https://doi.org/10.1016/J.ADDMA.2019.05.003.
ASTM D638-14 type i and type iv fused filament fabricated specimens, Polym. [120] N.G. Tanikella, B. Wittbrodt, J.M. Pearce, Tensile strength of commercial polymer
Test. 68 (2018) 294–301, https://doi.org/10.1016/J. materials for fused filament fabrication 3D printing, Addit. Manuf. 15 (2017)
POLYMERTESTING.2018.04.029. 40–47, https://doi.org/10.1016/J.ADDMA.2017.03.005.
[108] M. Shojib Hossain, D. Espalin, J. Ramos, M. Perez, R. Wicker, Improved [121] S.H. Masood, W. Rattanawong, P. Iovenitti, Part build orientations based on
mechanical properties of fused deposition modeling-manufactured parts through volumetric error in fused deposition modelling, Int J. Adv. Manuf. Technol. 16
(2000) 162–168.

22

You might also like