Elsevier

Polymer

Volume 53, Issue 25, 30 November 2012, Pages 5754-5759
Polymer

Top–down meets bottom–up: A comparison of the mechanical properties of melt electrospun and self-assembled 1,3,5-benzenetrisamide fibers

https://doi.org/10.1016/j.polymer.2012.10.016Get rights and content

Abstract

1,3,5-Benzenetrisamides (BTAs) are a prominent class of low-molecular weight compounds in supramolecular chemistry. They are well-known to self-assemble into micro- and nanofibers in a bottom–up approach. At the same time, BTAs are also suitable for top-down processing by melt electrospinning. In this work, we demonstrate for the first time that both approaches lead to mechanically robust BTA fibers. We compare self-assembled and electrospun fibers of N,N,N″-tripropyl-1,3,5-benzenetricarboxamide on multiple length scales. X-ray diffraction (XRD) reveals the same crystal structure independently from the preparation method. Using scanning electron microscopy (SEM), we observe significantly different morphologies of both fiber types on the sub-micron-scale. However, atomic force microscopy (AFM) bending experiments show that despite differences in morphology, Young's modulus is comparable for both types and in the lower GPa range (3–5 GPa). Thus, both top–down and bottom–up techniques with their complementary features in terms of accessible structures and potential applications are available for this class of materials.

Introduction

The controlled fabrication of well-defined microscopic fibrillar structures has become one of the main topics in materials science [1], [2], [3]. Networks and nonwovens based on these structures possess exceptional properties, such as high surface area, possibility for easy functionalization and superior mechanical strength [4]. These are promising features for applications such as tissue engineering, drug delivery, sensors, micro-/nanoelectromechanical systems (MEMS/NEMS), and filtration [3], [4], [5], [6], [7]. Especially fibers of sub-micron or nanoscale diameters are of interest due to their surface to volume ratio and the possibility to form structures with small mesh-sizes. Two approaches are feasible to access these length scales: Bottom–up approaches rely on the self-assembly of smaller units (even single molecules) to hierarchical structures [8], [9]. Top–down approaches, such as electrospinning, shape the materials directly into the desired structure. Especially for fibers and nonwovens, a great variety of structures has been demonstrated [10].

Both techniques are complementary in various ways: Since self-assembly allows simultaneous formation and growth of many fibers in a given volume, it is preferable in terms of processing times, especially for upscaling. In addition, if the processing conditions are chosen well, smaller fiber diameters are accessible in a more simple fashion than in electrospinning [9], [11]. On the downside, self-assembled fibers have smaller length and random orientation since they grow from many nuclei.

The advantage of electrospinning is that fibers can easily be formed with macroscopic length and well-defined orientation on macroscopic length scales. This even allows the controlled formation of superstructures at the micrometer level and above. However, the processing times are longer since electrospinning is a sequential process, in which the time necessary to form fibers is proportional to the total fiber length.

To offer the highest flexibility, it would be desirable to switch from one to the other approach for the same class of materials – especially for the formation of hierarchically organized structures which span multiple length scales.

Lately, the self-assembly of 1,3,5-benzenetrisamides (BTAs) into fibrillar structures has attracted increasing research interest [12], [13]. The benzene core realizes a planar and symmetric moiety and three amide groups allow the formation of strong hydrogen bonds between adjacent molecules resulting in supramolecular architectures [14]. BTAs are well-known as nucleating agents for polyvinylidene fluoride and polypropylene [15], [16], [17], [18]. Moreover, they are applied as organo- and hydrogelators [19], [20], [21], [22], as additives to improve the charge storage capability of electret materials [23] and as supramolecular materials [24], [25].

In addition to their bottom–up properties, we recently reported on the melt electrospinning of various BTAs and 1,3,5-cyclohexanetrisamides into defined fine fibers with a narrow size distribution [26]. Although a high molecular mass polymer is not essential for obtaining uniform electrospun fibers [27], electrospinning of low molecular weight substances is still unusual. BTAs form macrodipoles along the main axis of the column during the supramolecular assembly process within external electric fields and consequently offer excellent pre-conditions for electrospinning [28], [29], [30]. The melt electrospinning of BTAs is an exciting new top–down approach for self-assembling systems. It offers the possibility to overcome the strict limitation of the self-assembly conditions and consequently opens up a wide field of new applications for BTAs.

For all applications, a reasonable mechanical stability is an essential prerequisite. However, regardless by which means the BTA fibers are prepared, the mechanical characterization on a micron- or sub-micron scale requires sophisticated methods. A powerful technique is nanomechanical bending experiments, which have been used for the mechanical investigation of polymer nanofibers [31], [32], [33], biological materials [34], [35], [36], [37], [38], CNTs [39], [40], and nanowires [41], [42], [43]. In previous studies, we performed bending experiments on BTA micro- and nanofibers obtained via controlled self-assembly from nonpolar solvents. The experiments demonstrated that their molecular architecture allows control over the fiber morphology without decreasing their mechanical stability [44], [45].

In this work, we address the question whether the properties of BTA fibers are affected by using a top–down approach instead of a bottom–up approach. For that purpose, we prepare fibers of the same 1,3,5-benzenetrisamide via self-assembly from solution and melt electrospinning. This allows us for the first time to compare crystal structure, morphology and nano-mechanical properties of BTA fibers prepared from the same material.

Section snippets

Results and discussion

For the comparison, we prepared fibers of N,N′,N″-tripropyl-1,3,5-benzenetricarboxamide 1 (Scheme 1) via bottom–up (in the following termed SA-fibers) as well as top–down techniques (in the following termed ES-fibers). The SA-fibers were produced via controlled self-assembly by cooling of a solution of 1 in 2,2,4,4,6,8,8-heptamethylnonane (HMN). As top–down approach, we used melt electrospinning.

In order to investigate structural features on the Ångström-scale, we performed XRD measurements on

Conclusions

In this study we demonstrated for the first time that mechanically robust BTA fibers can be accessed via bottom–up- and top–down-approaches. We prepared self-assembled (SA) and melt electrospun (ES) fibers from the same compound 1 and obtained fibers with an average diameter of 1.2 ± 0.7 μm and 0.8 ± 0.2 μm for SA- and ES-fibers, respectively. On the Ångström-scale, XRD measurements show the same crystal structure of the fibers, independently of the preparation method. On the microscopic scale,

Material

N,N′,N″-tripropyl-1,3,5-benzenetricarboxamide 1 was synthesized according to the literature [50]. The melting temperature of 1 is 289 °C and was determined in a Perkin Elmer Diamond DSC (heating rate: 10 K/min, nitrogen flow: 20 mL/min). The temperature at a 10% weight loss is 351 °C. The measurement was performed in a Mettler SDTA 851 TGA at 10 K/min (nitrogen flow: 60 mL/min). Isothermal TGA runs at the spinning temperature (290 °C) under nitrogen atmosphere were performed to verify the

Acknowledgments

This work received financial support from the German Research Foundation (Deutsche Forschungsgemeinschaft) within the SFB 840, project B8. We thank Doris Hanft for help with the synthesis of the 1,3,5-benzenetrisamide, Martina Heider and Dr. Beate Förster from Bayreuth Institute of Macromolecular Research for support with the SEM images and Andreas Timme and Marina Behr for the XRD measurements. DK, JCS, BRN and JWN acknowledge the support of the Elite Network of Bavaria.

References (50)

  • S. Agarwal et al.

    Polymer

    (2008)
  • R.S. Barhate et al.

    Journal of Membrane Science

    (2007)
  • F. Abraham et al.

    Polymer

    (2010)
  • N. Mohmeyer et al.

    Polymer

    (2007)
  • A. Sakamoto et al.

    Polymer

    (2006)
  • C. Kulkarni et al.

    Chemical Physics Letters

    (2011)
  • L. Yang et al.

    Biomaterials

    (2008)
  • C. Guzman et al.

    Journal of Molecular Biology

    (2006)
  • D.R. Stamov et al.

    Biomaterials

    (2011)
  • Y.N. Xia et al.

    Advanced Materials

    (2003)
  • C. Kuttner et al.

    ACS Applied Materials & Interfaces

    (2012)
  • K. Jayaraman et al.

    Journal of Nanoscience and Nanotechnology

    (2004)
  • P. Moriarty

    Reports on Progress in Physics

    (2001)
  • M.P. Lutolf et al.

    Nature Biotechnology

    (2005)
  • L.C. Palmer et al.

    Accounts of Chemical Research

    (2008)
  • N. Kimizuka

    Self-assembly of supramolecular nanofibers

  • A. Greiner et al.

    Angewandte Chemie-International Edition

    (2007)
  • D.W. Hutmacher et al.

    Chemistry-an Asian Journal

    (2011)
  • I.A.W. Filot et al.

    The Journal of Physical Chemistry B

    (2010)
  • M. Kristiansen et al.

    Crystal Growth & Design

    (2009)
  • M.P. Lightfoot et al.

    Chemical Communications

    (1945–1946)
  • F. Abraham et al.

    Macromolecular Chemistry & Physics

    (2010)
  • M. Blomenhofer et al.

    Macromolecules

    (2005)
  • M. Schmidt et al.

    Crystal Growth & Design

    (2012)
  • Y. Yasuda et al.

    Chemistry Letters

    (1996)
  • Cited by (9)

    • Electrofluidodynamic technologies for biomaterials and medical devices: Melt electrospinning

      2018, Electrofluidodynamic Technologies (EFDTs) for Biomaterials and Medical Devices: Principles and Advances
    • Long-range interaction forces between 1,3,5-cyclohexanetrisamide fibers in crossed-cylinder geometry

      2016, Polymer
      Citation Excerpt :

      Under the influence of a strong electrical field, a thin jet is ejected from the so-called Taylor cone (see Fig. 1) [25]. Electrospinning of BTAs to homogenous [26] and mechanically stable [27] fibers with circular cross-section and smooth surfaces depends on various parameters such as molecular structure, temperature and viscosity of the melt, and the applied electric field. In this context, electrospinning of supramolecular fibers and avoiding the formation of solid spheres during the electrospinning process (i.e. electrospraying) is facilitated by using 1,3,5-cyclohexanetrisamides (CTAs) compared to BTAs [26,28].

    • Melt electrospinning today: An opportune time for an emerging polymer process

      2016, Progress in Polymer Science
      Citation Excerpt :

      Fig. 5 is a schematic summary of the different melt electrospinning configurations employed so far. One simple and often utilized method to supply polymer to a melt electrospinning system is to load polymer pellets into a syringe (either plastic [103,105,106,108,213,221,234–236], metal [230,237–239] or glass [31,200,205,206,217,223,240–243]) and preheat the material to remove any pockets of air so that an uninterrupted supply of polymer can be maintained. The syringe is then placed into a heating jacket to maintain a reservoir of polymer melt at a constant Tm.

    • Effects of hard and soft components on the structure formation, crystallization behavior and mechanical properties of electrospun poly(l-lactic acid) nanofibers

      2013, Polymer
      Citation Excerpt :

      Electrospinning [1–17] is a simple and versatile technique to produce polymer nanofibers with diameters ranging from tens of nanometers to micrometers.

    View all citing articles on Scopus
    1

    Tel.: +49 921 55 27 51; fax: +49 921 55 20 59.

    2

    Julia C. Singer and Daniel Kluge contributed equally to this work.

    View full text