Who invented honeycomb paper?

Author: Ingrid

Apr. 29, 2024

Packaging & Printing

Honeycomb structure - Wikipedia

Natural or man-made structures that have the geometry of a honeycomb

Read more Aluminum honeycomb structure Honeycomb structure in nature

Honeycomb structures are natural or man-made structures that have the geometry of a honeycomb to allow the minimization of the amount of used material to reach minimal weight and minimal material cost. The geometry of honeycomb structures can vary widely but the common feature of all such structures is an array of hollow cells formed between thin vertical walls. The cells are often columnar and hexagonal in shape. A honeycomb-shaped structure provides a material with minimal density and relative high out-of-plane compression properties and out-of-plane shear properties.[1]

Man-made honeycomb structural materials are commonly made by layering a honeycomb material between two thin layers that provide strength in tension. This forms a plate-like assembly. Honeycomb materials are widely used where flat or slightly curved surfaces are needed and their high specific strength is valuable. They are widely used in the aerospace industry for this reason, and honeycomb materials in aluminum, fibreglass and advanced composite materials have been featured in aircraft and rockets since the 1950s. They can also be found in many other fields, from packaging materials in the form of paper-based honeycomb cardboard, to sporting goods like skis and snowboards.

Introduction

[

edit

]

Natural honeycomb structures include beehives, honeycomb weathering in rocks, tripe, and bone.

Man-made honeycomb structures include sandwich-structured composites with honeycomb cores.[citation needed] Man-made honeycomb structures are manufactured by using a variety of different materials, depending on the intended application and required characteristics, from paper or thermoplastics, used for low strength and stiffness for low load applications, to high strength and stiffness for high performance applications, from aluminum or fiber reinforced plastics. The strength of laminated or sandwich panels depends on the size of the panel, facing material used and the number or density of the honeycomb cells within it. Honeycomb composites are used widely in many industries, from aerospace industries, automotive and furniture to packaging and logistics. The material takes its name from its visual resemblance to a bee's honeycomb – a hexagonal sheet structure.

History

[

edit

]

The hexagonal comb of the honey bee has been admired and wondered about from ancient times. The first man-made honeycomb, according to Greek mythology, is said to have been manufactured by Daedalus from gold by lost wax casting more than 3000 years ago.[2] Marcus Varro reports that the Greek geometers Euclid and Zenodorus found that the hexagon shape makes most efficient use of space and building materials. The interior ribbing and hidden chambers in the dome of the Pantheon in Rome is an early example of a honeycomb structure.[3]

Galileo Galilei discusses in 1638 the resistance of hollow solids: "Art, and nature even more, makes use of these in thousands of operations in which robustness is increased without adding weight, as is seen in the bones of birds and in many stalks that are light and very resistant to bending and breaking”.[4] Robert Hooke discovers in 1665 that the natural cellular structure of cork is similar to the hexagonal honeybee comb.[5] and Charles Darwin states in 1859 that "the comb of the hive-bee, as far as we can see, is absolutely perfect in economizing labour and wax”.[6]

The first paper honeycomb structures might have been made by the Chinese 2000 years ago for ornaments, but no reference for this has been found. Paper honeycombs and the expansion production process has been invented in Halle/Saale in Germany by Hans Heilbrun in 1901[7] for decorative applications. First honeycomb structures from corrugated metal sheets had been proposed for bee keeping in 1890.[8] For the same purpose, as foundation sheets to harvest more honey, a honeycomb moulding process using a paper paste glue mixture had been patented in 1878.[9] The three basic techniques for honeycomb production that are still used today—expansion, corrugation and moulding—were already developed by 1901 for non-sandwich applications.

Hugo Junkers first explored the idea of a honeycomb core within a laminate structure. He proposed and patented the first honeycomb cores for aircraft application in 1915.[10] He described in detail his concept to replace the fabric covered aircraft structures by metal sheets and reasoned that a metal sheet can also be loaded in compression if it is supported at very small intervals by arranging side by side a series of square or rectangular cells or triangular or hexagonal hollow bodies. The problem of bonding a continuous skin to cellular cores led Junkers later to the open corrugated structure, which could be riveted or welded together.

The first use of honeycomb structures for structural applications had been independently proposed for building application and published already in 1914.[11] In 1934 Edward G. Budd patented a welded steel honeycomb sandwich panel from corrugated metal sheets and Claude Dornier aimed 1937 to solve the core-skin bonding problem by rolling or pressing a skin which is in a plastic state into the core cell walls.[12] The first successful structural adhesive bonding of honeycomb sandwich structures was achieved by Norman de Bruyne of Aero Research Limited, who patented an adhesive with the right viscosity to form resin fillets on the honeycomb core in 1938.[13] The North American XB-70 Valkyrie made extensive use of stainless steel honeycomb panels using a brazing process they developed.

A summary of the important developments in the history of honeycomb technology is given below:[14]

  • 60 BC Diodorus Siculus reports a golden honeycomb manufactured by Daedalus via lost wax casting.
  • 36 BC Marcus Varro reports most efficient use of space and building materials by hexagonal shape.
  • 126 The Pantheon was rebuilt in Rome using a coffer structure, sunken panel in the shape of a square structure, to support its dome.
  • 1638 Galileo Galilei discusses hollow solids and their increase of resistance without adding weight.
  • 1665 Robert Hooke discovers that the natural cellular structure of cork is similar to the hexagonal honeybee comb.
  • 1859 Charles Darwin states that the comb of the hive-bee is absolutely perfect in economizing labour and wax.
  • 1877 F. H. Küstermann invents a honeycomb moulding process using a paper paste glue mixture.
  • 1890 Julius Steigel invents the honeycomb production process from corrugated metal sheets.
  • 1901 Hans Heilbrun invents the hexagonal paper honeycombs and the expansion production process.
  • 1914 R. Höfler and S. Renyi patent the first use of honeycomb structures for structural applications.
  • 1915 Hugo Junkers patents the first honeycomb cores for aircraft application.
  • 1931 George Thomson proposes to use decorative expended paper honeycombs for lightweight plasterboard panels.
  • 1934 Edward G. Budd patents welded steel honeycomb sandwich panel from corrugated metal sheets.
  • 1937 Claude Dornier patents a honeycomb sandwich panel with skins pressed in a plastic state into the core cell walls.
  • 1938 Norman de Bruyne patents the structural adhesive bonding of honeycomb sandwich structures.
  • 1941 John D. Lincoln proposes the use of expanded paper honeycombs for aircraft radomes
  • 1948 Roger Steele applies the expansion production process using fiber reinforced composite sheets.
  • 1969 Boeing 747 incorporates extensive fire-resistant honeycombs from Hexcel Composites using DuPont's Nomex aramid fiber paper.
  • 1980s Thermoplastic honeycombs produced by extrusion processes are introduced.

Manufacture

[

edit

]

Honeycomb crash absorption structure made of injection moulded thermoplastic polymer on a BMW i3

The three traditional honeycomb production techniques, expansion, corrugation, and moulding, were all developed by 1901 for non-sandwich applications. For decorative applications the expanded honeycomb production reached a remarkable degree of automation in the first decade of the 20th century.

Today honeycomb cores are manufactured via the expansion process and the corrugation process from composite materials such as glass-reinforced plastic (also known as fiberglass), carbon fiber reinforced plastic, Nomex aramide paper reinforced plastic, or from a metal (usually aluminum).[15]

Honeycombs from metals (like aluminum) are today produced by the expansion process. Continuous processes of folding honeycombs from a single aluminum sheet after cutting slits had been developed already around 1920.[16] Continuous in-line production of metal honeycomb can be done from metal rolls by cutting and bending.[17]

Thermoplastic honeycomb cores (usually from polypropylene) are usually made by extrusion processed via a block of extruded profiles[18] or extruded tubes[19][20] from which the honeycomb sheets are sliced.

Recently a new, unique process to produce thermoplastic honeycombs has been implemented, allowing a continuous production[21] of a honeycomb core as well as in-line production of honeycombs with direct lamination of skins into cost efficient sandwich panel.[22]

Applications

[

edit

]

Composite honeycomb structures have been used in numerous engineering and scientific applications.

More recent developments show that honeycomb structures are also advantageous in applications involving nanohole arrays in anodized alumina,[23] microporous arrays in polymer thin films,[24] activated carbon honeycombs,[25] and photonic band gap honeycomb structures.[26]

Aerodynamics

[

edit

]

Honeycombed, screened center for Langley's first wind tunnel

A honeycomb mesh is often used in aerodynamics to reduce or to create wind turbulence. It is also used to obtain a standard profile in a wind tunnel (temperature, flow speed). A major factor in choosing the right mesh is the length ratio (length vs honeycomb cell diameter) L/d.

Length ratio < 1: Honeycomb meshes of low length ratio can be used on vehicles front grille. Beside the aesthetic reasons, these meshes are used as screens to get a uniform profile and to reduce the intensity of turbulence.[27]

Length ratio >> 1: Honeycomb meshes of large length ratio reduce lateral turbulence and eddies of the flow. Early wind tunnels used them with no screens; unfortunately, this method introduced high turbulence intensity in the test section. Most modern tunnels use both honeycomb and screens.

While aluminium honeycombs are common use in the industry, other materials are offered for specific applications. People using metal structures should take care of removing burrs as they can introduce additional turbulences. Polycarbonate structures are a low-cost alternative.

The honeycombed, screened center of this open-circuit air intake for Langley's first wind tunnel ensured a steady, non-turbulent flow of air. Two mechanics pose near the entrance end of the actual tunnel, where air was pulled into the test section through a honeycomb arrangement to smooth the flow.

Honeycomb is not the only cross-section available in order to reduce eddies in an airflow. Square, rectangular, circular and hexagonal cross-sections are other choices available, although honeycomb is generally the preferred choice.[28]

Properties

[

Featured content:
Magnetic Closure Rigid Box Solutions at Unbeatable Prices
Why Plastic Office Storage Packaging is Essential
square food grade injection molding container

For more information, please visit Huanyou.

edit

]

A composite sandwich panel (A) with honeycomb core (C) and face sheets (B)

In combination with two skins applied on the honeycomb, the structure offers a sandwich panel with excellent rigidity at minimal weight. The behavior of the honeycomb structures is orthotropic, meaning the panels react differently depending on the orientation of the structure. It is therefore necessary to distinguish between the directions of symmetry, the so-called L and W-direction. The L-direction is the strongest and the stiffest direction. The weakest direction is at 60° from the L-direction (in the case of a regular hexagon) and the most compliant direction is the W-direction.[1] Another important property of honeycomb sandwich core is its compression strength. Due to the efficient hexagonal configuration, where walls support each other, compression strength of honeycomb cores is typically higher (at same weight) compared to other sandwich core structures such as, for instance, foam cores or corrugated cores.

The mechanical properties of honeycombs depend on its cell geometry, the properties of the material from which the honeycomb is constructed (often referred to as the solid), which include the Young's modulus, yield stress, and fracture stress of the material, and the relative density of the honeycomb (the density of the honeycomb normalized by that of the solid, ρ*/ρs).[29][30] The ratio of the effective elastic moduli and the solid's Young's moduli, e.g., κ ∗ / E s {\displaystyle \kappa ^{*}/E_{\text{s}}} and E ∗ / E s {\displaystyle E^{*}/E_{\text{s}}} , of low-density honeycombs are independent of the solid.[31] The mechanical properties of honeycombs will also vary based on the direction in which the load is applied.

In-plane loading: Under in-plane loading, it is often assumed that the wall thickness of the honeycomb is small compared to the length of the wall. For a regular honeycomb, the relative density is proportional to the wall thickness to wall length ratio (t/L) and the Young’s modulus is proportional to (t/L)3.[29][30] Under high enough compressive load, the honeycomb reaches a critical stress and fails due to one of the following mechanisms – elastic buckling, plastic yielding, or brittle crushing.[29] The mode of failure is dependent on the material of the solid which the honeycomb is made of. Elastic buckling of the cell walls is the mode of failure for elastomeric materials,[30] ductile materials fail due to plastic yielding, and brittle crushing is the mode of failure when the solid is brittle.[29][30] The elastic buckling stress is proportional to the relative density cubed, plastic collapse stress is proportional to relative density squared, and brittle crushing stress is proportional to relative density squared.[29][30] Following the critical stress and failure of the material, a plateau stress is observed in the material, in which increases in strain are observed while the stress of the honeycomb remains roughly constant.[30] Once a certain strain is reached, the material will begin to undergo densification as further compression pushes the cell walls together.[30]

Out of-plane loading: Under out-of-plane loading, the out-of-plane Young’s modulus of a regular hexagonal honeycombs is proportional to the relative density of the honeycomb.[29] The elastic buckling stress is proportional to (t/L)3 while the plastic buckling stress is proportional to (t/L)5/3.[29]

The shape of the honeycomb cell is often varied to meet different engineering applications. Shapes that are commonly used besides the regular hexagonal cell include triangular cells, square cells, and circular-cored hexagonal cells, and circular-cored square cells.[32] The relative densities of these cells will depend on their new geometry.

See also

[

edit

]

Honeycomb grille used on a computer fan to cover fan blades.

References

[

edit

]

Flexural Damage of Honeycomb Paperboard—A Numerical ...

This paper presents an experimental and numerical analysis using the finite element method (FEM) of the bending of honeycomb-core panel. Segments of honeycomb paperboard of several thicknesses were subjected to four-point flexure tests to determine their bending stiffness and maximum load. Several mechanical properties of orthotropic materials were taken into account to account for the experimental results. The numerical analysis of the damage prediction was conducted by using well-known failure criteria such as maximum stress, maximum strain and Tsai-Wu. The present study revealed how to model the honeycomb panel to obtain curves close to experimental ones. This approach can be useful for modelling more complex structures made of honeycomb paperboard. Moreover, thanks to the use of variously shaped cells in numerical models, i.e., the shape of a regular hexagon and models with a real shape of the core cell, results of the calculation were comparable with the results of the measurements. It turned out that the increase of maximum loads and rise in stiffness for studied samples were almost either linearly proportional or quadratically proportional as a function of the panel thickness, respectively. On the basis of failure criteria, slightly lower maximum loads were attained in a comparison to empiric maximum loads.

Based on the aforementioned literature one can see that there are a few papers related to strength/stiffness of honeycomb paperboards but the topic still isn’t exhausted. The present study concerns experimental and numerical analyses of honeycomb paperboard subjected to 4-point flexural tests. Experimental tests were performed for three thicknesses of honeycomb paperboard in the two main directions of paperboard plane, i.e., the machine direction (MD) and the cross direction (CD) to assess the influence on stiffness and the maximum load. Moreover, different cell shapes were modelled to reflect the real (imperfect) shapes of honeycombs. Simulations based on FEM were performed for five thicknesses of honeycomb paperboard and for two main directions (MD and CD). Furthermore, the authors of present study took into consideration three failure criteria to predict the damage of panels. Contrary to the present work, this approach wasn’t used in other works referring to modeling the honeycomb paperboard on purpose to determine the maximum loads. Nonlinear simulations computed in substeps were performed in ANSYS ® version 2019 R2 for large displacements based on the Green-Lagrange equations [ 48 ]. In addition, various mechanical properties of orthotropic materials were assumed to adjust the numerical characteristics to the experimental curves. Finally, assessment of different heights of panels and validation of numerical models including strength of paperboards can be useful to conduct further simulations on more complicated structures.

Honeycomb paperboards made from organic and biodegradable raw material are extensively used in numerous industries. Their main advantages are low specific weight, high strength and stiffness in relation to their specific weight [ 1 , 2 ]. Other advantages of honeycomb paperboard are excellent energy absorption properties, insulation, thermal and acoustic properties [ 3 ]. Honeycomb structures can be observed in Nature in bones, bees’ honeycombs or stalks of grain. Paper honeycomb and its expansion production process was invented in 1901 by Hans Heilbrun [ 4 ]. In the late decade of the 1930s Lincoln manufactured paper honeycomb from Kraft paper which subsequently was used in building furniture. The formed sandwich panel consisted of a thin hardwood facing bonded to a thick paper honeycomb core. Recently, paper honeycomb cores and paperboards can be found in everyday used objects. They can be found in many objects around us, starting from box endings, pallets, inserts, fillings, fillings for doors, furniture, partition walls in construction and sandwich multilayer structures used in the aviation and automotive industries [ 5 , 6 , 7 , 8 ]. Due to the usage of cheap and recyclable materials it is possible to reduce the price of final products as well as limit the usage of natural resources. In the literature, there are numerous studies concerning both experimental and numerical studies of paper products. The most popular studies concern the analysis of corrugated paperboards and cardboard boxes. Fadiji et al. [ 9 ] investigated the influence of the geometrical configuration of vents on the mechanical strength of packaging. Similar studies were developed in [ 10 , 11 , 12 ], where among others. Zaheer et al. [ 13 ] analysed the strength of paperboard packages subjected to compression loads by using the finite element analysis. Patel et al. [ 14 ] investigated the local buckling and collapse of corrugated boards under biaxial stress. The stability and collapse of the corrugated boards were analysed in [ 15 ]. Bai et al. [ 16 ] analysed axial crushing of single wall corrugated paperboards. Wang [ 17 ] investigated the cushioning properties of paper honeycomb under impact. Chen and Yan [ 18 ] performed a study which aimed at determining the elastic modulus of sandwich panels fitted with Kraft paper honeycomb. Gu et al. [ 19 ] researched the in-plane uniaxial crushing behavior of honeycomb paperboard by considering three types of deformation modes. Paperboards filled with honeycomb subjected to compression were analysed by Wang et al. [ 20 ]. The behavior of paper honeycomb panels subjected to bending and compression load was studied in [ 21 ], where a simplified model of a honeycomb core with small displacements was applied to determine only the stiffness. Analyses of the strength of paper tubes are presented in [ 22 , 23 , 24 ]. On other hand, works concerning analyses of composite structures’ stability from a theoretical approach can be found in [ 25 , 26 , 27 , 28 ] and works including additional experimental results are [ 29 , 30 , 31 , 32 , 33 , 34 , 35 , 36 , 37 , 38 ]. Bolzon and Talassi [ 39 ] investigated the behaviour of anisotropic paperboard composites till collapse by using burst strength testers. Mentrasti et al. examined the behaviour of creased paperboard experimentally [ 40 ] and analytically [ 41 ]. Borgqvist et al. [ 42 ] examined the continuum model of paperboard material with a high degree of anisotropy. Other papers devoted to studies of honeycomb cardboards are [ 43 , 44 , 45 , 46 ]. Hua et al. [ 43 ] investigated the influence on edgewise compressive strength of two sandwich paperboards by using numerical and experimental approaches. The authors of [ 44 ] studied honeycomb paperboards under impact compression by using FEM and experiments. Mou et al. [ 45 ] analysed the in-plane bearing capacity of a honeycomb paperboard based on plastic deformations and plastic energy dissipation. Other papers devoted to studies of honeycomb cardboards are collected in [ 46 , 47 ].

Orthotropic material was modelled as linearly elastic (i.e., obeying Hooke’s law). In the present paper, progressive damage for orthotropic material hasn’t been taken into account but the analysis of failure was based on the failure criteria which allow determining the initiation of damage and estimating the maximum load (low estimation of load-carrying capacity). In this research three failure criteria are used: maximal-stress, maximal-strain [ 48 ] and Tsai-Wu [ 51 ]. The tensile and compressive strength for both directions of orthotropy has to be determined experimentally. Due to the low thickness of paper, it is almost impossible to determine the shear strength of paper experimentally. Therefore, Equation (3) [ 14 ] was used to approximate shear strength of paper. The strength parameters of the paper are presented in .

In the FE model paper was modelled as linear-orthotropic in the elastic range [ 50 ]. The material properties were derived experimentally from data provided by the Lodz University of Technology Centre of Papermaking and Printing. presents the mechanical properties of paper used in FEA. Variation 1 corresponds to the nominal values of material properties. Variation 2 and 3 correspond to properties decreased or increased by 20% with respect to nominal values, respectively. The ±20-percent difference was used because it usually happens that the actual values of the mechanical properties of the papers can differ by up to 20% from the nominal values given in the specifications.

Numerical analysis was performed for a half of the panel (symmetry conditions were applied in the middle of the panel). The model was created by using shell 4node181 elements. According to the description [ 48 ], this element is suitable for analysing thin to moderately-thick shell structures. It is suitable for modelling composite shells or sandwich construction. While using shell elements it is necessary to assign thickness and material properties to the element. The walls of the honeycombs which are composed of two paperboards had double thickness (see and a). The process of preparation of the numerical model started from creating the honeycomb core. First, a single honeycomb cell was drawn by gluing neighbouring areas. It was performed by transforming the work plane coordinate system. Subsequently, areas making up single cell of honeycomb were copied and glued together to create the whole core of a sample of 200 mm × 100 mm dimensions. Lastly, faces were created and connected to the core. The face subjected to compression was modelled with the line in the place of support in order to ease application of boundary conditions. Due to small gaps between the copied geometry ( a) it was necessary to apply geometrical tolerances. The honeycomb core had seven elements along its height. The element edge length was set to 2 mm. Such attributes provided an overall good mesh quality. In order to create an appropriate model, it was necessary to modify the local coordinate system of single finite elements because the default orientations of the shell elements did not agree with the orthotropy directions. Four local coordinate systems were introduced and the element coordinate systems were transferred to correspond with the orthotropy directions as seen in the real model. Then shell normals were modified to keep the proper orientation of elements ( a). To simulate 4-point bending it was necessary to apply proper boundary conditions. At the support two translations were constrained. In the middle of the panel, symmetry boundary conditions were applied by constraining two rotations and one translation. The load was applied at the shorter edge of the model as the translation of all nodes at the upper edge of the structure. Applied boundary conditions are presented in b. The FE simulations were conducted for large displacements by using the Green-Lagrange formulation. The nonlinear computations performed in substeps were based on the Newton-Raphson algorithm. The number of substeps was set to be from 50 to 500. The number of iterations in each substep ranged from 10 to 500. The above described process was repeated for each of the examined paperboards. In numerical analysis paperboards with the following thicknesses were examined: Model 1—H = 8 mm, Model 2—H = 18 mm, Model 3—H = 22 mm, Model 4—H = 28 mm, and Model 5—H = 33 mm. In order to examine the influence of the honeycomb shape on the performance of the paperboard, it was necessary to create two models for each direction (MD and CD), one with the ideal honeycomb shape and second one with the real shape. Overall more than 60 numerical models were created.

Numerical analysis was performed for two shapes of honeycomb cell, because it was observed that real samples do not have a regular hexagon shape. a shows the shape of an ideal hexagonal core cell. b shows the modified shape of the core cell. presents the values of the geometric parameters of the ideal cell (regular hexagon) and the dimensions of the assumed cell. In the case of regular hexagon a = b and length of the regular hexagon’s side can be determined from Equation (2) using the parameter D of paperboards as delivered by the manufacturer ( ):

Initial forces of 3 N and 10 N was applied for H = 8 mm, H = 18 mm and H = 28 mm, respectively. The tests were performed until full failure of the specimens. During the analysis, the force vs. deflection in the middle of the structure was measured. In paper materials, the bending stiffness (BS) is one of the basic strength indicators and refers to the width of the bent sample. In the case of the 4−point bending tests, the BS was calculated according to Equation (1).

Before performing the bending tests, samples were dried at a temperature of 40 °C and subsequently they were conditioned according to standard PN-EN 20187:2000 (temperature 23 ± 1 °C and relative air humidity 50 ± 2%) [ 49 ]. The bending tests were carried out on Tensile Machine model Z010 (ZwickRoell, Ulm, Germany) equipped with a specialized tool ( a). The load range of the machine is from 0.1 N to 10 kN. The tool consists of four supports. Three of them (two upper ones and one lower) have two degrees of freedom (DoF) and the 4th one has only one DoF. During tests, the supports were moving at a velocity of 10 mm/min. The method used for placing the samples in the measuring grip is shown in b.

The distance between the supports and applied forces amounted to L 2 = 200 mm and 2L 1 + L 2 = 400 mm, respectively. The samples used in test had the surface dimensions of 100 mm × 500 mm. Due to the orthotopic properties of cardboard, the measurements were carried out in the two main directions (MD and CD). The method used for cutting samples is depicted in .

The experimental study was conducted in the Centre of Papermaking and Printing (Lodz University of Technology, Lodz, Poland). The research employed cellular cardboard in which both the core and the flat layers were made of Testliner type paper with a paperweight of 135 g/m 2 . Honeycomb paperboards with the same cell size (of D = 15 mm) and several thicknesses were tested. The following honeycomb paperboard thicknesses were considered: H = 8 mm, H = 18 mm and H = 28 mm. The 4−point bending tests of samples were carried out according to scheme depicted in .

The geometrical parameters of the cellular paperboard are described as: D —diameter of the circle inscribed in the regular hexagon (defined as the cell size), h —core height, H —paperboard thickness. Single thickness walls have the thickness of paper applied to manufacturing the cellular board core. However, walls glued to each other have double thicknesses.

Cellular cardboard is treated as an orthotropic body. This is due to two factors. First of all, it is caused by the nature of the core structure. Secondly, flat layers possess orthotropic material mechanical properties. In the plane of cellular paperboard, two main directions of orthotropy can be considered. The first one covers the direction of manufacturing (called machine direction or MD). The second (perpendicular) direction is called cross direction (CD). The main directions of the CD and MD of the paperboard are the same as paper used for the flat layers (CD O and MD O — ). In the case of a paperboard core, the machine direction of the paper applied for the MD R core is parallel to the height of the core. However, the cross direction CD R is perpendicular to the height of the core.

3. Results and Discussion

In this section, the research results are presented and discussed. The results obtained experimentally (bending stiffness, maximum load Fmax and force-deflection ratio ΔF/Δd in the maximum load range (from 10% to 50%) are given in .

Table 4

Denotation of the SampleMaximum Load
Fmax (N)BS (Nm)ΔF/Δd
(N/mm)Experiment 1.1 MD20.115.36.1Experiment 1.2 MD19.715.66.2Experiment 1.1 CD20.29.63.8Experiment 1.2 CD19.19.23.7Experiment 2.1 MD48.587.435.0Experiment 2.2 MD50.289.535.8Experiment 2.1 CD38.950.220.1Experiment 2.2 CD38.052.621.0Experiment 4.1 MD71.2179.471.8Experiment 4.1 CD65.3118.447.4Open in a separate window

The measurement results prove that as the thickness of the paperboard increases, the bending stiffness in both directions increases in the cardboard plane. In case of load-carrying capacity of samples under bending (in experiments), the maximum load rise corresponds almost proportionally to increase of paperboard thickness. It was also noticed that the determined bending stiffness (BS) and ΔF/Δd were quadratically proportional to the increase of the panel thickness. This is the result of an increase in the moment of inertia of the bent cross-section due to the increase in the sample thickness. By comparing the experimental results of BS and ΔF/Δd for MD and CD, a decrease by 35–45% was noted in the case of CD.

It has been observed that the location of the sample damage occurs in various places between the internal L2 supports. The reason of this effect is the presence of weakened places in the cardboard structure. Numerical models presented in this section are described according to the following scheme. Models with the close to the modified shape of the honeycomb cell contain letter R in the name of the model. The second number (number after the dot) describes the mechanical properties of the material (1—variation 1, 2—variation 2, 3—variation 3, see ). presents the curves of load vs. deflection for Model 1 (H = 8 mm) of paperboards for MD. It can be easily seen that numerical models are shown to be stiffer than the real structures. Surely, the shape of the honeycomb cell influences the stiffness of the numerical model. For small panel deflections, the shape of the considered cell doesn’t matter.

However, with the increase of the deflection, the stiffness of models with modifed honeycomb shape decreases significantly. The load-deflection curves for Model 1 (H = 8 mm) of paperboard bent in the CD are presented in .

In this case the differences in stiffness caused by the shape of the honeycomb cell are smaller than in the previous case. However, contrary to samples loaded in the MD, the models close to the real shape of the core tend to indicate slightly higher stiffness than ones with the ideal hexagonal core. Models with the second variation of mechanical properties behave like the structures in the experiments (curves Experiment 1.1 CD and Experiment 1.2 CD run close to Model 1.2 CD and Model 1.2 CD R). shows the load-deflection curves for the model with H = 18 mm in the MD. Like the 1st model in MD, the differences in stiffness caused by the change of shape of the honeycomb cell are small for a certain value of deflection. Then, an abrupt divergence of the stiffness is observed. It turns out that Model 2.2 MD R is the closest to the experimental results.

compares the load deflection curves obtained for both cell of honeycomb for model 2 in CD. The curves obtained for models with the 2nd variation of mechanical properties (as in the previous case) are the closest to the experimental data. The curves obtained for both honeycomb cell shapes are close to each other, but the models with the modified (close to real) honeycomb cell shape seem to be stiffer.

shows load deflection curve obtained for the 4th analysed paperboard in the MD (Model 4). As in previous analyses, the behaviour of the numerical model is not the same as that of the real object. In a certain range of deflections there is a small influence of the honeycomb cell shape on the stiffness of the structure. At a certain value of deflection, the stiffness changes significantly. Moreover, the curves obtained for models with reduced mechanical properties are closer to the experimental characteristics but the trends of these curves at the beginning are slightly different.

The load-deflection curves for model 4 in CD are presented in . The obtained characteristics of the load-deflection curves are similar to the previous ones. As it was noticed, load-deflection curves representing the behaviour of models with lowered mechanical properties are the closest to the experimental data. It is clear that models with a real honeycomb cell shape show higher stiffness, while models with the ideal honeycomb cell shape are closer to the experimental curves.

As presented in , , , , and in each case the load-deflection curves attained for models with variation 2 of mechanical properties seem to be closest to the experimental study results. It should also be noted that the model 2 in the direction of CD reflects very well the relationship between the force F and the deflection d obtained by experimental study. This means that it can be used to calculate the bending stiffness of cardboard, which is determined by the value of the ΔF/Δd ratio (see Equation (1)). Therefore, the results presented for another variants are limited to variation 2 (with reduced mechanical properties). displays the comparison of charts for all considered models in the MD. The stiffness of the paperboard increases with the increase of the thickness of the paperboard. The influence of the shape of the honeycomb cell is insignificant in a certain range of deflections, but at a certain point the influence of the shape of the honeycomb cell became significant. Models with modified honeycomb cell shape appear to be less stiff.

compares the load-deflection curves for models in the CD. The general observations are similar to ones for models in the MD. However, in this case a reverse situation was observed because models with real honeycomb cell shape seemed to be stiffer than models with ideal the honeycomb cell shape.

The location of the failure propagation (place at which first signs of failure are observed) and the corresponding loads obtained for three failure criteria (Max-Stress denotes maximum stress criterion, Max-Strain represents the maximum strain criterion and Tsai-Wu means simply the Tsai-Wu criterion) are presented in . It can be easily seen that the failure load depends on the thickness of the paperboard. Moreover, the shape of the honeycomb cell has an essential impact on the results. The propagation of failure in models with ideal honeycomb cell shape is observed at higher loads than in models with modified honeycomb cell shape. The differences in the failure loads for models in the MD for different failure criteria are rather insignificant. In all cases, the first signs of failure (where stress states in these points were fulfilled based on the failure criterion) occurred in the outer layer subjected to compression (it happened mostly at the supports). After comparing the data presented in with , and it is visible that the failure forces attained for models with an ideal honeycomb cell shape are closer to the experimental study results. Moreover, the failure location indicated by the Tsai-Wu criterion corresponds to that observed in the experiments. presents the comparison of failure locations obtained for the Tsai-Wu criterion with the experimental results. In the case of numerical models, only half of the panel is presented. a,b show that the failure occurred in the region near the support, while c indicates the region near the center of modelled panel.

Table 5

ModelCell
TypeTsai-WuMax-StressMax-StrainFailure
Load
[N]Failure LocationFailure Load
[N]Failure
LocationFailure
Load
[N]Failure
LocationModel 1.2MD19Near support19Near support20In the middleMD R15In the middle15In the middle15In the middleModel 2.2MD44Near support44Near support44Near supportMD R35In the middle35In the middle35In the middleModel 3.2MD54Near support55In the middle55In the middleMD R41In the middle41In the middle41In the middleModel 4.2MD72In the middle72In the middle72In the middleMD R51In the middle51In the middle51In the middleModel 5.2MD84In the middle84In the middle84In the middleMD R62In the middle62In the middle62In the middleOpen in a separate window

In , the failure loads based on the considered failure criteria are given. Referring to the failure criteria, the differences in maximum loads are comparable. However, by comparing the maximum loads based on numerical results and experimental ones, close correlations were observed (19 N for Model 1.2 MD and 15 N in the case of Model 1.2 MD R vs. 20.1 N for Experiment 1.1 MD and 19.7 N for Experiment 1.2 MD). In case of Model 2 MD, 44 N for Model 2.2 MD and 35 N for Model 2.2 MD R vs. 48.5 N for Experiment 2.1 MD, 50.2 N for Experiment 2.2 MD was noticed. In the case of Model 4 MD 72 N for Model 4.2 MD and 51 N for Model 4.2 MD R in a comparison to 71.2 N for Experiment 4.1 MD and 50.2 N for Experiment 4.2 MD was noted. presents the failure loads and locations obtained for the investigated models in the CD. Contrary to the MD, the propagation of failure for panels in the CD is seen at higher loads for models with real honeycomb cell shapes. Comparison of the data presented in with , and leads to the conclusion that the failure loads obtained in the numerical analysis are smaller than the experimental ones. The failure loads obtained for models with real honeycomb cell shapes are closer to the ones obtained in the experimental study. In a comparison of failure loads determined numerically and experimentally for the CD, the following results were attained: 12 N for Model 1.2 CD and 13 N in case of Model 1.2 CD R vs. 20.2 N for Experiment 1.1 CD and 19.1 N for Experiment 1.2 CD. In the case of Model 2 for CD, the failure loads are the following: 26 N for Model 2.2 CD and 33 N for Model 2.2 CD R vs. 38.9 N for Experiment 2.1 CD, 38.0 N for Experiment 2.2 CD. In the case of Model 4 CD, 48 N for Model 4.2 CD and 59 N for Model 4.2 CD R were registered, in contrast to 65.3 N for Experiment 4.1 CD. In general, the numerical failure loads seemed to be smaller but it should be mentioned that the failure criteria allow indicating the initiation of failure (often called low estimation of load-carrying capacity) but not an accurate moment of the damage.

Table 6

ModelCell TypeTsai-WuMax-StressMax-StrainFailure
LoadFailure LocationFailure LoadFailure
LocationFAILURE
LoadFailure
LocationModel 1.2CD12Near support12Near support12Near supportCD R13In the middle13In the middle14In the middleModel 2.2CD26Near support26Near support28Near supportCD R33In the middle33In the middle33In the middleModel 3.2CD32Near support34Near support34Near supportCD R40In the middle40In the middle42In the middleModel 4.2CD48In the middle48In the middle48In the middleCD R59Near support64Near support64Near supportModel 5.2CD49Near support49Near support49Near supportCD R67In the middle67In the middle69In the middleOpen in a separate window

The Tsai-Wu criterion provides the lowest failure load for a particular cell type among all analysed failure criteria. illustrates the failure locations obtained for numerical models with real honeycomb cell shapes for the Tsai-Wu criterion with the experimental data. The failure locations indicated in a,b are in agreement with the data presented in . In the case of c, reaching the maximum stresses based on the given failure criterion (on the scale index greater than 1) was observed in the region near the support, but then the maximal value of the Tsai-Wu index was observed in the middle of the panel like in case of the experimental study.

presents maps of the failure index on the surface of the face subjected to compression for the three failure criteria and six consecutive load steps (forces from 5 N up to 30 N). A value on the scale smaller than 1 defines the safe region before the failure occurs. A value on the scale (index) greater than 1 means the limit stress state based on some failure criterion was exceeded. In these locations, damage propagations might develop. The greater value of the index on the scale means a greater possibility of local failure of the material. Based on maps, it can be observed that extremes of the Tsai-Wu and Max-Stress indexes are present in the support region but at some load value the extremes move from the support regions to the middle of the panel. It is worth mentioning that the transfer occurs at a force slightly higher than the force at which the ultimate stress state occurred (i.e., where the failure index is greater than 1). In the case of the Max-Strain criterion the failure location does not change. The failure index maps for the Tsai-Wu criterion differ slightly from the failure index maps for the Max-Stress and Max-Strain criteria because the Tsai-Wu criterion also contains in its formula [51] a factor representing the shear stress.

Table 7

F = 5 NF = 10 NF = 15 NF = 20 NF = 25 NF = 30 NTsai-Wu Max-Stress Max-Strain Open in a separate window

The distribution of failure indexes on faces subjected to compression for the three analysed failure criteria and five consecutive load steps for Model 1.2 CD R are presented in (forces from 5 N up to 25 N). It can be observed that the Max-Stress criterion and Max-Strain criterion indicate almost the same locations of failure which for all steps are in regions close to the middle of the panel. In case of the Tsai-Wu failure index, its extreme is observed in the middle of the panel, but then it changes its position twice from the center of the panel to the region of the support, and then again to the middle of the panel.

Table 8

F = 5 NF = 10 NF = 15 NF = 20 NF = 25 NTsai-Wu Max-Stress Max-Strain Open in a separate window

The behavior of panels presented in and shows that in the case of real specimens the failure location should be expected to appear in multiple locations. Due to manufacturing issues, local damage during handling or sample preparation might happen so a real specimen will not fail in the place indicated as the one in which propagation of failure should begin. If the real specimen, due to the mentioned issues, happens to be stronger in the place at which it should fail first, or weaker in the place which is an extremum of the failure criteria at higher force (as presented in and ) then failure will occur in other locations. Based on the comparison of the results of measurements and calculations of the destructive load of cardboard in the MD bending test, it can be stated that the results of calculations using Model 2 do not differ much from the results obtained in the experiments.

If you want to learn more, please visit our website Honeycomb Paper Manufacturers.

14

0

Comments

Please Join Us to post.

0/2000

All Comments ( 0 )

Guest Posts

If you are interested in sending in a Guest Blogger Submission,welcome to write for us!

Your Name: (required)

Your Email: (required)

Subject:

Your Message: (required)