Comparison of compressive properties of layered syntactic foams having gradient in microballoon...

11
1 Citation Information: Material Science and Engineering A. Vol. 427, No. 1-2, 2006, pp. 331-342. Comparison of Compressive Properties of Layered Syntactic Foams having Gradient in Microballoon Volume Fraction and Wall Thickness Nikhil Gupta and William Ricci Composite Materials and Mechanics Laboratory Mechanical, Aerospace and Manufacturing Engineering Department Polytechnic University Brooklyn, NY 11201 ABSTRACT The existing functionally graded syntactic foams (FGSFs) are based on creating a gradient of microballoon (hollow particles) volume fraction along the length or thickness of the composite to achieve a variation in density and mechanical properties. However, such an approach has several limitations. Variation in volume fractions of the constituents leads to the possibility of warping or localized swelling of FGSF if it is exposed to varying temperature and moisture conditions. It can also lead to premature fracture due to nonuniform distribution of stress along the microballoon gradient and existence of resin/microballoon rich sides in the material. Hence, a FGSF material needs to be created that is independent of the volume fraction variation. The present study attempts to address this requirement. A FGSF is fabricated that is based on the microballoon wall thickness variation along the specimen length. In this novel approach the volume fraction is available as an additional parameter for controlling the properties of the FGSF. The FGSFs based on volume fraction gradient (VF-type) and on microballoon wall thickness gradient (RR-type) are fabricated in a layered structure, and evaluated for compressive properties. It is observed that the VF-type FGSFs show a sharp drop in stress, on the order of 40-60%, after the peak compressive strength value. However, such feature is not observed in RR-type FGSFs leading to the possibility of gaining better control over strength and energy absorption. The compression of such foams could be continued to 60-75% strain based on the microballoon volume fraction in their structure. The total energy absorption was found to be 3-5 times higher in RR- type FGSFs compared to VF-type FGSFs and plain syntactic foams. Keywords: Syntactic foams, functionally graded material, microballoon, mechanical properties, porosity. INTRODUCTION The closed cell foams can have either single-phase or two-phase microstructures. The microstructure of single- phase closed cell foams contains pores which are not interconnected [1, 2]. Structural features of these foams provide them with higher strength compared to the open cell foams, which contain interconnected pores [3]. Two-phase closed cell foams, called syntactic foams, are made by dispersing low density hollow particles (microballoons) in a matrix material [4]. The microstructure of syntactic foams is shown in Figure 1, Figure 1. A scanning electron micrograph showing the structure of a syntactic foam. Glass microballoons are embedded in an epoxy resin matrix. where glass microballoons are dispersed in an epoxy resin matrix. The presence of strong but lightweight particles leads syntactic foams to higher strength compared to single- phase closed cell foams. Almost all types of foams show high energy absorption, damage tolerance and vibration damping capabilities. Structure and Properties of Syntactic Foams It is possible to tailor properties of syntactic foams by changing a wide variety of parameters, which include the matrix and microballoon materials, and microballoon size, size distribution, wall thickness and volume fraction [5-7]. Quasi- static and dynamic properties of syntactic foams can be found studied in great detail in the published literature [8-15]. Syntactic foams are currently used in marine and aerospace applications because of their low density, which is attributed to the presence of low density microballoons in their structure. Increase in microballoon volume fraction (V mb ) decreases the density of syntactic foams. Hence, most applications tend to use syntactic foams having V mb in the range of 0.3-0.65. In this V mb range most syntactic foams show a stress-plateau in their compressive stress-strain curve, which leads to high energy absorption [5, 16]. The stress- plateau is observed in metal matrix syntactic foams also [17, 18]. The presence and extent of the plateau region is greatly affected by the microballoon wall thickness, among other parameters [19].

Transcript of Comparison of compressive properties of layered syntactic foams having gradient in microballoon...

1

Citation Information: Material Science and Engineering A. Vol. 427, No. 1-2, 2006, pp. 331-342.

Comparison of Compressive Properties of Layered Syntactic Foams having Gradient in

Microballoon Volume Fraction and Wall Thickness

Nikhil Gupta and William Ricci

Composite Materials and Mechanics Laboratory

Mechanical, Aerospace and Manufacturing Engineering Department

Polytechnic University

Brooklyn, NY 11201

ABSTRACT

The existing functionally graded syntactic foams

(FGSFs) are based on creating a gradient of microballoon

(hollow particles) volume fraction along the length or

thickness of the composite to achieve a variation in density

and mechanical properties. However, such an approach has

several limitations. Variation in volume fractions of the

constituents leads to the possibility of warping or localized

swelling of FGSF if it is exposed to varying temperature and

moisture conditions. It can also lead to premature fracture due

to nonuniform distribution of stress along the microballoon

gradient and existence of resin/microballoon rich sides in the

material. Hence, a FGSF material needs to be created that is

independent of the volume fraction variation. The present

study attempts to address this requirement. A FGSF is

fabricated that is based on the microballoon wall thickness

variation along the specimen length. In this novel approach

the volume fraction is available as an additional parameter for

controlling the properties of the FGSF. The FGSFs based on

volume fraction gradient (VF-type) and on microballoon wall

thickness gradient (RR-type) are fabricated in a layered

structure, and evaluated for compressive properties. It is

observed that the VF-type FGSFs show a sharp drop in stress,

on the order of 40-60%, after the peak compressive strength

value. However, such feature is not observed in RR-type

FGSFs leading to the possibility of gaining better control over

strength and energy absorption. The compression of such

foams could be continued to 60-75% strain based on the

microballoon volume fraction in their structure. The total

energy absorption was found to be 3-5 times higher in RR-

type FGSFs compared to VF-type FGSFs and plain syntactic

foams.

Keywords: Syntactic foams, functionally graded material,

microballoon, mechanical properties, porosity.

INTRODUCTION

The closed cell foams can have either single-phase or

two-phase microstructures. The microstructure of single-

phase closed cell foams contains pores which are not

interconnected [1, 2]. Structural features of these foams

provide them with higher strength compared to the open cell

foams, which contain interconnected pores [3]. Two-phase

closed cell foams, called syntactic foams, are made by

dispersing low density hollow particles (microballoons) in a

matrix material [4]. The microstructure of syntactic foams is

shown in Figure 1,

Figure 1. A scanning electron micrograph showing the structure of a

syntactic foam. Glass microballoons are embedded in an epoxy resin

matrix.

where glass microballoons are dispersed in an epoxy

resin matrix. The presence of strong but lightweight particles

leads syntactic foams to higher strength compared to single-

phase closed cell foams. Almost all types of foams show high

energy absorption, damage tolerance and vibration damping

capabilities.

Structure and Properties of Syntactic Foams

It is possible to tailor properties of syntactic foams by

changing a wide variety of parameters, which include the

matrix and microballoon materials, and microballoon size, size

distribution, wall thickness and volume fraction [5-7]. Quasi-

static and dynamic properties of syntactic foams can be found

studied in great detail in the published literature [8-15].

Syntactic foams are currently used in marine and

aerospace applications because of their low density, which is

attributed to the presence of low density microballoons in their

structure. Increase in microballoon volume fraction (Vmb)

decreases the density of syntactic foams. Hence, most

applications tend to use syntactic foams having Vmb in the

range of 0.3-0.65. In this Vmb range most syntactic foams

show a stress-plateau in their compressive stress-strain curve,

which leads to high energy absorption [5, 16]. The stress-

plateau is observed in metal matrix syntactic foams also [17,

18]. The presence and extent of the plateau region is greatly

affected by the microballoon wall thickness, among other

parameters [19].

2

Higher density syntactic foams show higher strength

and modulus compared to lower density foams as shown in

Figure 2 [20, 21]. However, they have two distinct

disadvantages, namely lower fracture strain leading to lower

energy absorption and a 20-40% decrease in stress

immediately after attaining the peak compressive strength.

While, the compressive strength and the plateau stress are

among the design parameters and are specified by the

application, the total plateau strain can be increased in order to

increase the energy absorption by the material in the process

of compression.

Figure 2. Compressive stress-strain curves for plain syntactic foams

having different densities.

Fiber reinforcement of syntactic foams has been

studied as a means of increasing their strength [22-25].

Several studies have focused on increasing the energy

absorption in higher density syntactic foams by means of

incorporating second phase particles such as nanoclay and

rubber [26, 27]. However, moisture absorption was found to

increase considerably in some hybrid syntactic foams, which

could be detrimental in both marine and aerospace structures

[28].

Functionally Graded Syntactic Foams

In syntactic foams, an approach of obtaining high

energy absorption under compression may involve fabricating

them in functionally graded microstructure, which is explored

in the present study. Several studies are available on

functionally graded syntactic foams (FGSFs) [29-33]. In these

studies a gradient in Vmb is created along one dimension of

the material as schematically shown in Figure 3. These

FGSFs are referred here as VF-type FGSFs. Initiation and

propagation of cracks in the resin rich side of the specimen

under quasi-static and dynamic loading conditions appear to

be the problems with these kinds of FGSFs.

Creation of Vmb gradient severely limits the

possibility of tailoring the physical and mechanical properties

of syntactic foam to the desired level. Normally, the matrix

has higher thermal expansion and moisture absorption

coefficients compared to that of microballoons. Hence, high

matrix volume fraction on one side of the FGSF causes

nonuniform thermal expansion or moisture absorption when

the material is exposed to changing temperature or moisture

conditions, which can lead to warping and localized swelling.

Among other limitations, variation in Vmb causes nonuniform

stress concentrations in the material, which may cause

catastrophic failure, especially when the matrix is a brittle

polymer such as epoxies. Additionally, this approach poses

severe limitations on the minimum density achievable in such

FGSFs because Vmb cannot be varied independently. Hence,

an approach independent of Vmb variation is required for the

fabrication of FGSFs.

Figure 3. The schematic representation of the existing syntactic foam

microstructure based on microballoon volume fraction variation.

A FGSF is studied in the present research that can

have Vmb constant in the material microstructure but still have

a gradient in mechanical properties by means of changing

microballoon wall thickness. Previous studies have observed

that the strength and modulus of syntactic foams increase with

increased wall thickness of microballoons used in synthesizing

them [21, 34]. Hence, a gradient in mechanical properties of

syntactic foams can be obtained if wall thickness is varied

across the material thickness as schematically shown in Figure

4 [35]. These types of FGSFs are referred here as the RR-type.

Having the same Vmb in all parts of the structure substantially

reduces the limitations of variation in coefficient of thermal

expansion, moisture absorption coefficient and the fracture

mode of the FGSFs. Microscopic observations cannot provide

evidence for the presence of functionally graded structure in

RR-type FGSFs because the only parameter that varies is the

internal radius of microballoons (or the wall thickness). As

observed in Figure 5, two types of syntactic foams containing

microballoons of about the same average outer diameter but

different inner diameters look nearly the same because of the

same Vmb (0.3). Figure 5a and 5b show microstructures of

syntactic foams, named SF220 and SF460, containing

microballoons of wall thickness of about 1 and 2.5 m,

respectively. However, the density difference between various

layers of FGSFs can provide clear evidence of having a graded

structure, which will be explored in this study.

Specimen Failure Pattern

It is found that the microballoon material, matrix

material, microballoon wall thickness and Vmb are important

factors in deciding the mechanical properties and fracture

mode of syntactic foams. Three fracture modalities can exist

in syntactic foams based on experimental and theoretical

studies, which are reviewed by Bardella and Genna [19].

Incre

asin

g m

icro

ball

oo

n

vo

lum

e f

racti

on

Matrix

Microballoons

0

10

20

30

40

50

60

70

80

0 0.1 0.2 0.3 0.4 0.5

Strain (mm/mm)

Str

es

s (

MP

a)

493 kg/m3

650 kg/m3

3

Figure 4. The schematic representation of functionally graded

syntactic foam microstructure based on variation in microballoon .

(a)

(b)

Figure 5. Scanning electron micrographs of (a) SF220 and (b) SF460

syntactic foams containing 30% microballoons by volume.

The first two, longitudinal crack initiation in the

direction of applied load and shear failure at 45 , have been

widely observed and documented by various researchers in

experimental and theoretical investigations [6, 21, 34]. Metal

matrix syntactic foams were also found to fail due to the

failure of microballoons along the shear directions [18, 36].

High wall thickness of the microballoons and high volume

fraction of rigid matrix material are found to be the parameters

that lead to the longitudinal cracking in the material. The third

modality is observed by only a few researchers [11, 36, 37],

which constitutes origination of failure along some specific

weak planes in transverse direction. A review of Kim’s

experimental scheme reveals that their syntactic foam

fabrication method is based on settling of microballoons in a

fluid. Such methods have less probability of providing

random distribution of microballoons in the syntactic foams. It

can create planes of microballoons having the same wall

thickness and strength leading to generation of some weak

planes. These weak planes can be selectively crushed during

compression. Hence, the failure mechanism of such materials

is different from the first two mechanisms.

It is also observed that at high strain rates the first

two fracture modalities prevail [12, 38]. As the strain rate

increases, the material has lesser time to respond to the applied

load and longitudinal cracks become the primary fracture

mode. The crack initiation and growth is closely monitored in

the present study and is compared to these existing

observations.

POROSITY ASPECTS IN SYNTACTIC FOAMS

It is important to develop an understanding of the

porosity existing inside the syntactic foam microstructure

because it significantly affects the material properties,

especially the densification limit of foams. Two types of

porosities exist in syntactic foams, termed as microballoon and

matrix porosities, which are schematically illustrated in Figure

6. The microballoon porosity is enclosed within the

microballoons and is the desired type of porosity. This

porosity can be controlled by selecting microballoons of

appropriate size and wall thickness and using them in desired

volume fraction. During the synthesis of syntactic foams

some air gets entrapped within the matrix between

microballoons, as illustrated in Figure 6, which is termed as

the matrix porosity. The matrix porosity is normally

undesirable as it reduces the strength of the composite and can

lead to increased moisture absorption.

Figure 6. Schematic representation of the syntactic foam

structure showing microballoon and matrix porosities

The porosity plays an important role in determining

physical and mechanical properties of syntactic foams. A

detailed analysis of the effect of porosity on the mechanical

properties of syntactic foams is presented by Bardella and

Genna [6, 19]. During compression testing the compressing

material fills up the matrix porosity.

Microballoon

Porosity Matrix

Porosity

Matrix

Microballoons

Incre

asin

g m

icro

ball

oo

n w

all

thic

kn

ess

(d

ecre

as

ing

)

Matrix

Microballoons

4

As microballoons fracture the porosity enclosed

within them is exposed, which can also be occupied by the

compressing material. Hence, the length of the stress-plateau

in the stress-stain curve of syntactic foams is directly

dependent on the total porosity in the material structure,

among other factors. The calculation method for both types of

porosities is outlined here, which will be used in the

estimation of total porosity in FGSFs.

The wall thickness of microballoons is defined by

means of a parameter named Radius Ratio ( ), which is given

by Equation 1 [34]:

o

i

r

r (1)

where ri and ro are the inner and outer radii of microballoons.

The higher values of lead to lower wall thicknesses (w) in

the same size microballoons, as given by Equation 2.

1rw o (2)

In the bulk testing it is observed that the strength of

microballoons increases with increased wall thickness or a

decrease in . Generally, for ceramic particles the

compressive strength of an individual particle increases as the

wall thickness decreases [39]. Such a trend is observed

because ceramics are very susceptible to fracture due to flaws.

Microballoons have a variety of flaws embedded in their walls

[40]. However, the bulk behavior of microballoons is opposite

to this trend and microballoons with high wall thickness show

higher strength. This trend is observed because in the bulk

testing the microballoon fracture takes place primarily due to

the shear and not due to the true compression.

The hollow fraction of the microballoon volume (Vc,mb) can

be given as 3

,mbcV (4)

Hence, the volume fraction of microballoon porosity in

syntactic foam structure is 3

mbc VV (5)

The difference between the theoretical ( th) and

experimentally measured ( m) densities of syntactic foams

can be used to get an estimate of the matrix porosity (Vv) [41].

th

mthvV (6)

where the theoretical density is calculated using Rule of

Mixtures:

rmbmbmbth V1V (7)

where r and mb are the densities of the resin system and

microballoons, respectively. The total porosity in syntactic

foams (Vt) is the sum of the microballoon and the matrix

porosities:

vct VVV (8)

During synthesis of syntactic foams it is inevitable to

cause damage to some of the microballoons. Careful

processing can minimize the damage but cannot completely

eliminate it. Fracture of microballoons opens up their cavity,

which can be filled with the matrix material. Hence, fracture

of microballoons leads to increase in the syntactic foam

density and reduces Vc. Therefore, the calculation of

microballoon and matrix porosities cannot be very precise, but

can only be approximate.

MATERIALS AND EXPERIMENTAL METHODS

Constituent Materials

DER 332 epoxy resin with hardener DEH 24,

manufactured by DOW Chemical Co., is used as the matrix

resin in the fabricated syntactic foams. A diluent C12-C14

aliphaticglycidylether is used in 5 vol.% quantity to reduce the

viscosity of the resin and facilitate mixing of microballoons in

high volume fractions. Twenty types of plain syntactic foams

are synthesized with four types of microballoons in five

volume fractions of 0.3, 0.4, 0.5, 0.6 and 0.65. Scotchlite

glass microballoons, manufactured and supplied by 3M, are

used in synthesizing these foams. The outer diameter, density

and calculated for the selected microballoons are given in

Table 1. A difference in their densities is caused by the

values because their average outer diameters are nearly the

same. The compressive strength and modulus for various types

of plain syntactic foams synthesized in this study are presented

in Table 2. These values are measured experimentally from

the synthesized syntactic foams.

Table 1. Properties of microballoons used in synthesizing syntactic

foams.

Layer

#

Syntactic

Foam

Type

Average

Micro-

balloon

Size

( m)

True

Particle

Density

(kg/m3)

Pressure

for Min.

80%

Fractional

Survival

(MPa)

Radius

Ratio

1 SF220 35 220 2.8 0.9703

2 SF320 40 320 13.8 0.9561

3 SF380 40 380 27.6 0.9474

4 SF460 40 460 41.4 0.9356

Table 2. Nomenclature and compressive properties of plain syntactic

foam layers used in fabricating functionally graded syntactic foams.

Micro-

balloon

Density

(kg/m3)

Micro-

balloons

(Vol.%)

Syntactic

Foam

Nomenclature

Strength

(MPa)

Modulus

(MPa)

SF220 30 SF220-30 67.9 1966

40 SF220-40 56.6 1803

50 SF220-50 42.9 1512

60 SF220-60 34.6 1429

65 SF220-65 22.6 895

SF320 30 SF320-30 87.5 2282

40 SF320-40 80.7 2252

50 SF320-50 66.7 2052

60 SF320-60 51.0 1878

65 SF320-65 37.7 1406

SF380 30 SF380-30 102.2 2326

40 SF380-40 91.7 2351

50 SF380-50 66.8 2087

60 SF380-60 62.6 2099

65 SF380-65 50.4 1785

SF460 30 SF460-30 105.2 2508

40 SF460-40 100.8 2414

50 SF460-50 89.5 2473

60 SF460-60 70.4 2260

65 SF460-65 56.1 1844

5

Fabrication of Syntactic Foams

The fabrication process consists of mixing resin,

diluent and hardener and stirring them until a uniform mixture

is obtained. Microballoons are then added in desired volume

fraction. The processing is carried out at 500 C to reduce the

viscosity of the matrix resin system, which helps in achieving

uniform mixing of microballoons in the resin system. The

mixture is gently stirred using wooden dowels to obtain a

slurry, which is degassed for 10 minutes and then cast in steel

molds. The cast slabs are cured at 250 C for 24 h and post

cured at 1000 C for 3 h.

Preparation of FGSF Specimens and Compression Testing

Foam chips of 12 12 mm2 cross section area and 5

mm thickness were sectioned and bonded together with thin

layers of an epoxy based adhesive to create VF- and RR-type

FGSFs. In VF-type FGSFs, chips of syntactic foams having

the same type of microballoons in Vmb of 0.3-0.6 were bonded

together as schematically shown in Figure 7a. To fabricate the

RR-type FGSFs, foam chips having different types of

microballoons in the same Vmb were bonded together as an

example shown in Figure 7b. Four compositions of VF-type

and five compositions of RR-type FGSF were fabricated as

per the details given in Tables 3 and 4, respectively.

(a)

(b)

Figure 7. Specimen configuration of the functionally graded syntactic

foam tested for compressive properties (a) VF-type and (b) RR-type.

Each layer has 5 mm thickness in these specimens.

Four specimens of each type were prepared and

tested under compression using a computer controlled Instron

4467 material test system. The crosshead velocity was

maintained at 1 mm/min for the compression tests. The load

and displacement data was recorded and used for developing

the stress-strain curves.

Table 3. Density of VF-type functionally graded syntactic foams.

FGSF

Type

Foam

Type

Layer

#

Theoretical

Density

(kg/m3)

Measured

Density

(kg/m3)

Measured

Density

of FGSF

(kg/m3)

VF220 SF220-30 1 878 850

SF220-40 2 784 735

SF220-50 3 690 611

SF220-60 4 596 550 686

VF320 SF320-30 1 908 875

SF320-40 2 824 780

SF320-50 3 740 670

SF320-60 4 656 611 734

VF380 SF380-30 1 926 887

SF380-40 2 848 800

SF380-50 3 770 694

SF380-60 4 692 625 751

VF460 SF460-30 1 950 931

SF460-40 2 880 846

SF460-50 3 810 732

SF460-60 4 740 645 788

Table 4. Density of RR-type functionally graded syntactic foams.

FGSF

Type

Foam

Type

Layer

#

Theoretical

Density

(kg/m3)

Measured

Density

(kg/m3)

Measured

Density of

FGSF

(kg/m3)

RR30 SF220-30 1 878 850

SF320-30 2 908 875

SF380-30 3 926 887

SF460-30 4 950 931 885

RR40 SF220-40 1 784 735

SF320-40 2 824 780

SF380-40 3 848 800

SF460-40 4 880 846 790

RR50 SF220-50 1 690 611

SF320-50 2 740 670

SF380-50 3 770 694

SF460-50 4 810 732 677

RR60 SF220-60 1 596 550

SF320-60 2 656 611

SF380-60 3 692 625

SF460-60 4 740 645 608

RR65 SF220-65 1 549 504

SF320-65 2 614 543

SF380-65 3 653 588

SF460-65 4 705 654 572

The modulus values are calculated as the slope of the

linear portion of the stress-strain curves between the strain of

1 and 2.5%. Such a range of strain was selected because some

of the microballoons have very low strength and may fracture

at higher deformations. The number of microballoons

fractured may not be sufficient to cause any visible shift in the

linear trend. However, use of data in the selected strain range

reduces any possible uncertainties. Similar range is used for

the calculation of modulus of RR-type FGSFs also.

SF220-60

SF320-60

SF380-60

SF460-60

Layer 1

Layer 2

Layer 3

Layer 4

12 mm

12 mm

SF220-30

SF220-40

SF220-50

SF220-60

Layer 1

Layer 2

Layer 3

Layer 4

12 mm

12 mm

6

RESULTS AND DISCUSSION

Porosity Calculation

The total porosity (Vt) is one of the important

parameters that define the energy absorption capabilities of

syntactic foams. Hence, Vt is calculated for each layer of

FGSFs using Equations 5, 6 and 8. Vt affects the length of the

stress-plateau in the compressive stress-strain curves of

syntactic foams, in addition to several other parameters such

as properties of matrix material and matrix microballoon

interfacial bonding. However, in the present case the only

differences between various types of foams are the

microballoon wall thickness and volume fraction (both leading

to difference in porosity). Hence, the effect of only these two

parameters on the energy absorption in syntactic foams is

discussed.

It is possible to calculate the microballoon porosity

by theoretical means using Equation 5; however, calculation

of matrix porosity requires theoretical estimation and

experimental measurement of foam densities. The theoretical

density is calculated using Equation 7. The experimental

density is calculated using ASTM C271 standard method,

where at least 5 pieces of foam were measured for weight and

volume to calculate the density [42]. The calculated and

measured density values are given in Tables 3 and 4. These

values are inserted in Equation 6 to calculate the matrix

porosity, and in Equation 8 to calculate the total porosity. The

porosity values of all types of syntactic foams are presented in

Tables 5 and 6. These values are used to calculate the densities

of the FGSF specimens.

Table 5. Total porosity in VF-type functionally graded syntactic foam

specimens.

FGSF

Type

Layer # Micro-

balloon

Porosity

(Vol.%)

Matrix

Porosity

(Vol.%)

Total

Porosity

in Each

Layer

(Vol.%)

Porosity

in the

FGSF

(Vol.%)

VF220 1 27.4 3.2 30.6

2 36.5 6.3 42.8

3 45.7 11.4 56.7

4 54.8 7.7 62.8 48.2

VF320 1 26.2 3.6 29.8

2 35.0 5.3 40.3

3 43.7 9.5 52.7

4 52.4 6.9 59.4 45.5

VF380 1 25.5 4.2 29.7

2 34.0 5.7 39.7

3 42.5 9.9 52.5

4 51.0 9.7 61.0 45.7

VF460 1 24.6 2.0 26.6

2 32.8 3.9 36.7

3 40.9 9.6 50.9

4 49.1 12.8 61.1 43.8

The measured densities and total porosities in all

variations of VF- and RR-type syntactic foams are also

presented in Tables 3-6. It can be observed from Table 5 that

the average Vt in all VF-type FGSFs is around 0.45.

Microballoon does not have any significant effect on the

porosity, which is controlled only by the Vmb value. All

varieties of VF-type FGSFs have variation in Vmb over the

range 0.3-0.6. To cause an effective increase in porosity the

volume fraction range of microballoons will have to be

increased.

Table 6. Total porosity in RR-type functionally graded syntactic foam

specimens.

FGSF

Type

Layer

#

Micro-

balloon

Porosity

(Vol.%)

Matrix

Porosity

(Vol.%)

Total

Porosity

in Each

Layer

(Vol.%)

Mean

Porosity

in the

FGSF

(Vol.%)

RR30 1 27.4 3.2 30.6

2 26.2 3.6 29.8

3 25.5 4.2 29.7

4 24.6 2.0 26.6 29.2

RR40 1 36.5 6.3 42.8

2 35.0 5.3 40.3

3 34.0 5.7 39.7

4 32.8 3.9 36.7 39.9

RR50 1 45.7 11.4 56.7

2 43.7 9.5 52.7

3 42.5 9.9 52.5

4 40.9 9.6 50.9 53.2

RR60 1 54.8 7.7 62.8

2 52.4 6.9 59.4

3 51.0 9.7 61.0

4 49.1 12.8 61.1 61.1

RR65 1 59.4 8.2 67.4

2 56.8 11.6 67.8

3 55.3 10.0 65.3

4 53.2 7.2 60.2 65.2

The RR-type FGSFs show a clear trend in Vt. The

specimens containing 30 and 65 vol.% microballoons have Vt

of 0.29 and 0.65, respectively. Hence, the variation in the

porosity can be achieved by selecting microballoons of

appropriate values. Having higher Vmb of high

microballoons will provide lower density and higher Vt in

composites.

Close observations of Tables 5 and 6 reveal that Vt in

various layers of VF-type specimens varies from about 0.29 to

0.6, whereas in RR-type specimens Vt in each layer is within a

close range. The syntactic foams with lower porosity show

higher strength. Hence, a difference in the compressive

properties of VF-type and RR-type FGSFs is expected because

of a difference in pattern of porosities in the specimen

microstructure.

Compression Tests

FGSFs Based on Volume Fraction Variation (VF-Type)

The representative compressive stress-strain curves

for two specimens of each VF-types FGSFs are shown in

Figure 8. It can be noted that except for VF220, other

compositions of VF-type FGSFs show a sharp drop in stress,

40-60%, immediately after achieving the peak compressive

strength. Such a drop is enough to be termed as complete

failure in a large number of applications. The drop

corresponds to initiation of a longitudinal crack (in the

direction of compression), as shown in Figure 9. Such cracks

are common occurrence in the compression testing of plain

syntactic foams containing low microballoons, such as

SF380 and SF460, and play a significant role in the final

failure of the specimen [34, 43].

7

The initiation of longitudinal cracks has been widely

related to the microballoon and Vmb, along with other

possible parameters such as matrix properties and specimen

aspect ratio which are not varied in the present study [19]. It

can be noticed in Figure 9 that the longitudinal crack is cutting

across various layers of syntactic foams. Hence, the material

fracture behavior is not affected by the layered structure in any

significant manner. As observed in Figure 10 the crack

growth has taken place primarily in the matrix material.

Microballoon fracture is not significant in this figure, which

resembles the tensile failure [44]. In these specimens the resin

rich side is prone to crack initiation under the secondary

tensile stresses.

0

10

20

30

40

50

60

70

0 0.2 0.4 0.6 0.8

Strain (mm/mm)

Str

ess (

MP

a)

(a)

0

10

20

30

40

50

60

70

0 0.2 0.4 0.6 0.8

Strain (mm/mm)

Str

ess (

MP

a)

(b)

0

10

20

30

40

50

60

70

0 0.2 0.4 0.6 0.8Strain (mm/mm)

Str

ess (

MP

a)

(c)

0

10

20

30

40

50

60

70

0 0.2 0.4 0.6 0.8

Strain (mm/mm)

Str

es

s (

MP

a)

(d)

Figure 8. Representative stress-strain curves for (a) VF220, (b)

VF320, (c) VF380 and (d) VF460-type functionally graded syntactic

foams having volume fraction gradient. Each graph shows curves for

two specimens of the same type to demonstrate their repeatability

characteristics.

Figure 9: Specimen failure mode for VF460 type FGSFs. The

longitudinal crack can be observed in the specimen.

Figure 10: SEM of the fracture surface generated due to the

longitudinal crack shown in Figure 9 in SF460-40 layer.

VF460-30

VF460-40

VF460-50

VF460-60

8

It can be observed from Figure 8 that the stress-strain

curves stabilize after the drop in stress and show a stress-

plateau. The plateau region continues till the densification is

complete. In most of the VF-type FGSFs the porosity volume

is around 45% but the densification appears to be completed

around 60% strain because the standard test conditions and

specimen geometry allow for some lateral expansion, leading

to appearance of additional strain.

VF220-type FGSFs did not show drop in the stress

after the peak compressive strength. Compression testing of

SF220 plain syntactic foams also showed similar trend. The

high microballoons used in SF220 foams start fracturing

early in the compression testing due to their significantly

lower strength compared to that of the matrix resin. Hence, the

fracture mode of SF220 plain syntactic foams and VF220

FGSFs is found to be different from corresponding syntactic

foams having other types of microballoons. A compressed

VF220-type FGSF is shown in Figure 11, where the high Vmb

layers can be observed crushed because of compression and

the low Vmb layers have a longitudinal crack. The crack

initiation was delayed in these specimens and most energy in

the initial stages of compression was absorbed in the

densification of weak layers.

Figure 11: Specimen failure mode for VF-220 type FGSFs. SF220-60

layer is almost completely crushed and SF220-50 layer is partially

crushed by the time the longitudinal crack propagated.

The peak compressive strength and modulus values

of the VF-type syntactic foams are listed in Table 7. It can be

observed that for an increase in the FGSF density from 686 to

788 kg/m3 the compressive strength and modulus increase by

about 108% and 19% respectively. In plain syntactic foams

also the strength and modulus are found to increase with

density. The specific strength and modulus for all VF-type

FGSFs are around 0.068 and 2.1 MPa/kg/m3. However, the

corresponding values for various compositions of plain

syntactic foams are 0.09 and 2.8 MPa/kg/m3. The specific

strength and modulus of VF-type FGSFs are lower because the

modulus and the peak strength are defined by those of the

weakest layer in the FGSF, whereas the density is increased by

the presence of heavier compositions. In VF-type FGSFs the

plateau stress is found to be around 30 MPa and the energy

absorption during compression, calculated as the area under

the stress-strain curve, is found to be in the range of 14-18

MPa, which is significantly higher than 3-9 MPa [26]

calculated for most plain syntactic foams.

Table 7. Compressive properties of VF-type functionally graded

syntactic foams.

FGSF Type Compressive Modulus

(MPa)

Peak Compressive

Strength

(MPa)

VF220 1362 30.2

VF320 1562 54.6

VF380 1886 51.1

VF460 1620 62.7

Observations of the fracture mode and stress-strain

curves reveal that VF-type syntactic foams have higher energy

absorption compared to the plain syntactic foams. However,

the specific strength and modulus of plain syntactic foams are

higher because of their relatively lower density. Hence, VF-

type FGSFs present only limited advantage compared to plain

syntactic foams under compressive loading conditions.

FGSFs Based on Radius Ratio Variation

The RR-type syntactic foams contain low and high

microballoons in the same specimen. In all compositions of

RR-type FGSFs the SF220 layer starts fracturing early in the

compression, delaying the crack initiation and slowing down

its propagation. Hence, sharp drop in stress is not observed in

the stress-strain curves for these specimens shown in Figure

12. In low Vmb (0.3 and 0.4) specimens longitudinal cracking

is observed due to high stiffness of the specimens [19, 21].

However, the crack initiation was delayed and led to decrease

in stress in a progressive manner instead of a sharp drop and

the lowest stress is observed consistently around 40% strain.

In higher Vmb specimens the longitudinal cracking is

suppressed until the SF220 layer is completely crushed as

shown in Figure 13, compared to the VF-type FGSFs shown in

Figure 9. The SEM micrograph presented in Figure 14 shows

that microballoons are completely crushed in SF220 layer.

Only debris and small fragments can be observed in the

microstructure. It is observed from the stress-strain curves

that as the Vmb is increased the stress-plateau becomes more

stable. For FGSFs containing 50-65 vol.% microballoons there

is no effective decrease in stress after the peak compressive

strength and the stress-plateau is observed to appear at the

same stress level.

0

10

20

30

40

50

60

0 0.15 0.3 0.45Strain (mm/mm)

Str

ess (

MP

a)

(a)

SF220-30

SF220-40

SF220-50

SF220-60

9

0

10

20

30

40

50

60

0 0.15 0.3 0.45 0.6 0.75

Strain (mm/mm)

Str

ess (

MP

a)

(b)

0

10

20

30

40

50

60

0 0.15 0.3 0.45 0.6 0.75Strain (mm/mm)

Str

ess (

MP

a)

(c)

0

10

20

30

40

50

60

0 0.15 0.3 0.45 0.6 0.75Strain (mm/mm)

Str

ess (

MP

a)

(d)

0

10

20

30

40

50

60

0 0.15 0.3 0.45 0.6 0.75

Strain (mm/mm)

Str

es

s (

MP

a)

(e)

Figure 12. Compressive stress-strain curves for (a) RR30, (b) RR40,

(c) RR50, (d) RR60 and (e) RR65-type functionally graded syntactic

foams containing microballoon wall thickness gradient. Each graph

shows curves for two specimens of the same type to demonstrate

their repeatability characteristics.

(a)

(b)

(c)

(d)

Figure 13: Specimen failure mode for RR60-type FGSF. (a) the

specimen before compression (b) the SF220 layer (top) starts to crush

10

(c) the SF220 layer is half crushed and cracks can be seen in SF320

layer (d) SF220 layer has crushed completely and in other layers

longitudinal cracks can be seen, which are not of sufficient length to

cause complete failure.

Figure 14: SEM of the failed SF220 layer observed in Figure 13

showing complete crushing of microballoons.

Table 8. Compressive properties of RR-type functionally graded

syntactic foams.

FGSF Type Compressive

Modulus

(MPa)

Peak Compressive

Strength

(MPa)

RR30 1582 59.4

RR40 1314 52.1

RR50 1366 49.0

RR60 1411 35.0

RR65 1060 27.7

The compressive modulus and strength values for

RR-type FGSFs are presented in Table 8. Similar to that of the

VF-type FGSFs, the strength and modulus of RR-type FGSFs

are close to the corresponding values for plain syntactic foams

of SF220 type that have the same Vmb. The modulus and

strength of the FGSF depend on those of the weakest layer

used in their structure, which is SF220 in the present case.

Once the densification of the weakest layer is completed the

stress starts rising and the next layer starts deforming, which

sometimes causes appearance of another peak in the stress-

strain curve. When the complete densification is achieved the

stress-plateau region ends and the stress starts increasing. The

strain for complete densification depends on the total porosity

in the foam specimens, among other factors, which varies

between 29 and 65 vol.% for FGSFs containing 30 – 65 vol.%

microballoons, as given in Table 6. The average plateau

stresses for the FGSFs containing 0.65, 0.6 and 0.5 Vmb are

found to be 23, 29 and 37 MPa. It can be noticed that the

stress is constant within a narrow band in the plateau region

and there is no visible sign of failure of the specimen in the

stress-strain curves. Hence, the use of high Vmb RR-type

FGSFs in structural applications in may enhance the safety

and reliability.

To obtain an estimate of total energy absorbed during

compression the area under the stress-strain curves is

measured to be 16, 16.5 and 22 MPa for FGSFs specimens

containing Vmb of 0.65, 0.6 and 0.5, respectively. Similar

calculations for specimens containing Vmb of 0.3 and 0.4

result in energy absorption of 15 and 18 MPa. The energy

absorption is the highest in specimens having Vmb of 0.5

because of higher strength of these specimens. The values of

energy absorption in RR-type FGSFs are higher than those for

plain syntactic foams and for VF-type FGSFs.

The approach of changing the microballoon , while

keeping the Vmb constant, has shown higher compressive

failure strain leading to higher energy absorption. The new

FGSF microstructure provides precise control over the

porosity in the material structure and desired level of strength

and modulus. Several parameters exist in FGSFs that can lead

to design of materials with predetermined strength, modulus

and energy absorption. The arrangement of layers in a desired

order can lead to structures that fail at a known location as per

the design requirements. The better ability to tailor the

properties of these FGSFs and higher energy absorption will

help in increasing the safety standards in the structures where

these materials are used.

CONCLUSIONS

Functionally graded syntactic foams (FGSFs) are

studied for compressive loading conditions in the present

study. The structure is based on variation in the wall thickness

of microballoons instead of the more common approach of

volume fraction variation. The FGSFs based on volume

fraction variation have shown sudden drop of 40-60% in the

stress immediately after the peak compressive strength. Such

foams showed limited enhancement in compressive properties

compared to the plain syntactic foams. The undesirable drop

in the stress is not observed in FGSFs that are based on the

new approach of creating the microballoon wall thickness

gradient. The total energy absorption was found to be 200-

500% more in these FGSFs than the comparable plain

syntactic foams. The replacement of plain syntactic foams

with new FGSFs in structural applications can significantly

enhance the structural safety under compressive loading. The

compressive strength and modulus of FGSFs can be increased

by choosing microballoons of lower values. The new

approach is independent of variation in microballoon volume

fraction. Hence, volume fraction can also be changed to

synthesize foams of desired strength and modulus.

ACKNOWLEDGMENTS

Funding for the research was provided by Othmer

Institute at Polytechnic University under the major project

initiation grant, which is gratefully acknowledged. 3M is

acknowledged for providing glass microballoons and technical

information related to them. Help of Tien Chih Lin and

Sandeep Kumar Gupta in parts of the experimental program is

also acknowledged.

REFERENCES

1. N.J. Mills, H.X. Zhu, J. Mech. Phys. Solid. 47, 3 (1999)

669-695.

2. K.Y.G. McCullough, N.A. Fleck, M.F. Ashby, Acta

Mater. 47, 8 (1999) 2323-2330.

3. M.N. Bureau, R. Gendron, J. Cell. Plast. 39 (2003) 353-

367.

11

4. F.A. Shutov, in: D. Klempner, K.C. Frisch (Eds.),

Handbook of Polymeric Foams and Foam Technology,

second ed., Hanser Publishers, New York, 2004, Ch. 17.

5. P. Bunn, J.T. Mottram, Composites. 24 (1993) 565-571.

6. L. Bardella, F. Genna, Int. J. Sol. Struct. 38, 40-41 (2001)

7235-7260.

7. C.S. Karthikeyan, S. Sankaran, Kishore, Mater. Lett. 58, 6

(2004) 995-999.

8. J.S. Huang, L.J. Gibson, J. Mech. Phys. Solid., 41, 1

(1993) 55-75.

9. S.J. Park, F.L. Jin, C. Lee, Mater. Sci. Eng. A. 402, 1-2

(2005) 335-340.

10. N. Gupta, R. Nagorny, J. Appl. Polym. Sci. 2006,

Accepted, in press.

11. H.S. Kim, P. Plubrai, Composites Part A. 35, 9 (2004)

1009-1015.

12. D. Rittel, Mater. Lett. 59, 14-15 (2005) 1845-1848.

13. D. Benderly, Y. Rezek, J. Zafran, D. Gorni, Polym.

Compos. 25 (2004) 229-236.

14. B. Song, W. Chen, T. Yanagita, D.J. Frew, Compos.

Struct. 67, 3 (2005) 279-287.

15. E. Rizzi, E. Papa, A. Corigliano, Int. J. Sol. Struct. 37, 40

(2000) 5773-5794.

16. N. Gupta, Kishore, E. Woldesenbet, S. Sankaran, J.

Mater. Sci. 36 (2001) 4485-4491.

17. P.K. Rohatgi, J.K. Kim, N. Gupta, S. Alaraj and A.

Daoud. Composites Part A, 37, 3 (2006) 430-437.

18. D.K. Balch, J.G. O’Dwyer, G.R. Davis, C.M. Cady, G.T.

Gray III, D.C. Dunand, Mater. Sci. Eng. A. 391, 1-2,

2005, 408-417.

19. L. Bardella, F. Genna, in: F.K. Ko, G.R. Palmese, Y.

Gogotsi, A.S.D. Wang (Eds.), Proceedings of 20th

Annual

Technical Conference of the American Society for

Composites, Philadelphia, PA, USA, DEStech

Publications, Lancaster, PA, 2005, paper #101.

20. E.M. Wouterson, F.Y.C. Boey, X. Hu, S.C. Wong,

Compos. Sci. Technol. 65 (2005) 1840-1850.

21. N. Gupta, E. Woldesenbet, Composites Part A. 35 (2004)

103-111.

22. Y.J. Huang, L. Vaikhanski and S.R. Nutt, Composites

Part A. 37, 3 (2006) 488-496.

23. M.V. Alonso, M.L. Auad, S.R. Nutt, Compos. Sci.

Technol. Corrected Proof, Available online 14 February

2006, doi:10.1016/j.compscitech.2005.12.016.

24. M.V. Alonso, M.L. Auad, S. Nutt, Composites Part A.

Corrected Proof, Available online 7 February 2006,

doi:10.1016/j.compositesa.2005.12.011.

25. C.S. Karthikeyan, S. Sankaran, M.N. Jagdish Kumar,

Kishore, J. Appl. Polym. Sci. 81, 2 (2001) 405-411.

26. N. Gupta, R. Maharsia, H.D. Jerro. Mater. Sci. Eng. A.

395 (2005) 233-240.

27. N. Gupta, R. Maharsia, Appl. Compos. Mater. 12 (2005)

247-261.

28. R. Maharsia, N. Gupta, H.D. Jerro, in: F.K. Ko, G.R.

Palmese, Y. Gogotsi, A.S.D. Wang (Eds.), Proceedings of

20th

Annual Technical Conference of the American

Society for Composites, Philadelphia, PA, USA, DEStech

Publications, Lancaster, PA, 2005, paper # 79.

29. Kishore, R. Shankar, S. Sankaran, Mater. Sci. Eng. A.

412, 1-2 (2005) 153-158.

30. V. Parameswaran, A. Shukla, J. Mater. Sci. 35 (2000) 21-

29.

31. V. Parameswaran, A. Shukla, Mech. Res. Commun. 29

(2002) 397-405.

32. M.A. El-Hadek, H.V. Tippur, Int. J. Sol. Struct. 40 (2003)

1885-1906.

33. C.E. Rousseau, H.V. Tippur, Eng. Fract. Mech. 69 (2002)

1679-1693.

34. N. Gupta, E. Woldesenbet, J. Cell. Plast. 40 (2004) 461-

480.

35. N. Gupta. Mater. Lett. Submitted, 2005.

36. M. Kiser, M.Y. He, F.W. Zok, Acta Mater. 47, 9 (1999)

2685-2694.

37. J.A. DeRuntz Jr., O. Hoffman, J Appl. Mech. 31 (1969)

551-557.

38. E. Woldesenbet, N. Gupta, A. Jadhav, J. Mater. Sci. 40,

15 (2005) 4009-4017.

39. M. Koopman, G. Gouadec, K. Carlisle, K.K. Chawla, G.

Gladysz. Scripta Mater. 50, 5 (2004) 593-596.

40. N. Gupta, in: N. Gupta and W. H. Hunt (Eds.),

Proceedings of Solidification Processing of Metal Matrix

Composites – The Rohatgi Honorary Symposium, TMS

Annual Meeting 2006, March 12-16, San Antonio, TX,

TMS Publications, Warrendale, PA, pp. 3-14.

41. B.D. Agarwal, L.J. Broutman, Analysis and Performance

of Fiber Composites, second ed., John Wiley, New York,

1990.

42. C271-99 Standard Test Method for Density of Sandwich

Core Materials. ASTM International, West

Conshohocken, PA, USA.

43. B. Song, W. Chen, D.J. Frew, J. Compos. Mater. 38

(2004) 915-936.

44. N. Gupta, E. Woldesenbet, Kishore, J. Mater. Sci., 37, 15

(2002) 3199-3209.