Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2333 2023-04-17 10:01:25 |
2 layout & references Meta information modification 2333 2023-04-18 03:24:03 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Bandara, S.; Wijesundara, K.; Rajeev, P. Ultra-High-Performance Fibre-Reinforced Concrete. Encyclopedia. Available online: https://encyclopedia.pub/entry/43110 (accessed on 20 May 2024).
Bandara S, Wijesundara K, Rajeev P. Ultra-High-Performance Fibre-Reinforced Concrete. Encyclopedia. Available at: https://encyclopedia.pub/entry/43110. Accessed May 20, 2024.
Bandara, Sahan, Kushan Wijesundara, Pat Rajeev. "Ultra-High-Performance Fibre-Reinforced Concrete" Encyclopedia, https://encyclopedia.pub/entry/43110 (accessed May 20, 2024).
Bandara, S., Wijesundara, K., & Rajeev, P. (2023, April 17). Ultra-High-Performance Fibre-Reinforced Concrete. In Encyclopedia. https://encyclopedia.pub/entry/43110
Bandara, Sahan, et al. "Ultra-High-Performance Fibre-Reinforced Concrete." Encyclopedia. Web. 17 April, 2023.
Ultra-High-Performance Fibre-Reinforced Concrete
Edit

Ultra-high-performance fibre-reinforced concrete (UHPFRC) is a cementitious composite which contains fibres. UHPFRC has emerged as an effective structural retrofitting material due to its superior mechanical properties. In addition, UHPFRC has outstanding durability, ductility and workability; a low permeability; and a high abrasion and fire resistance. 

ultra-high-performance fibre-reinforced concrete rehabilitation strengthening durability

1. Ultra-High-Performance Fibre-Reinforced Concrete (UHPFRC)

UHPFRC is one of the revolutionary discoveries in cement- and concrete-related composites. It can be identified as a cementitious composite material comprised of steel fibres as a reinforcement, replacing conventional reinforcing steel. The initial development of UHPFRC began in the 1970s by investigating high-strength cement pastes with lower water/cement ratios. These pastes were complemented with fibres, superplasticizers and pozzolanic admixtures and paved the way for the introduction of UHPFRC [1]. UHPFRC possesses superior mechanical properties of a compressive strength exceeding 150 MPa, a direct tensile strength higher than 7–8 MPa and a flexural strength more than 30 MPa [2][3][4]. In addition, UHPFRC exhibits outstanding durability, ductility and workability; a low permeability; and a high abrasion resistance, fire resistance and impact strength [5]. These improved characteristics are obtained by replacing the coarse aggregate with a well-graded fine aggregate to obtain a homogeneous concrete matrix, introducing high-strength steel fibres to improve the ductile behaviour, lowering the water/binder ratio and introducing a super-plasticizer and a highly active pozzolanic material [4][6].
The main differences between UHPFRC and typical reinforced concrete are the presence of the fibres, the size of the aggregate and the amount of binder. The matrix of UHPFRC is much denser in comparison with conventional concrete and the improved mechanical and durability properties are attributed to dense matrix. Achieving the maximum packing density of the granular components is paramount in producing UHPFRC [7]. The introduction of steel fibres enhances the tensile and flexural strength of UHPFRC. Moreover, the steel fibres have the potential to absorb tensile stresses, mitigating the spread and linkage of microcracks [8]. Steel fibres in the UHPFRC mix can have different aspect ratios. It is recommended to use 2–4% steel fibres by mixture volume for a workable and economical UHPFRC mix [9]. The mix proportions of UHPFRC should be selected considering both the economic and sustainable aspects in achieving a denser mix and the optimum composition to gain improved mechanical properties and durability. Different particle packing models have been investigated by researchers to obtain an optimum packing in the matrix and hence to achieve a denser mix [10][11]. A typical mix design of UHPFRC contains Portland cement, silica fume, crushed quartz, fine sand, superplasticizer, steel fibres and water. Most of the commercially available UHPFRC products have been developed in European countries such as France and Germany. In addition, Japan and USA also have succeeded in manufacturing commercial UHPFRC products. Table 1 illustrates the typical mix proportion range by weight in UHPFRCs and the exact mix proportions in some of the commercially available UHPFRC products. The typical chemical compositions (percentage mass) of ordinary Portland cement (OPC) and other supplementary cementitious materials used for UHPFRC mix can be found in Wang et al. [12], and thus are not presented here. 
Table 1. UHPFRC typical mix design proportion range and mix proportions of commercial UHPFRC products [Adapted from Refs. [13][14]].

2. Mechanical Properties of UHPFRC

2.1. Compressive Strength

According to the standards of material testing, compressive strength can be measured either on cubes or cylinders, provided that the conversion factor between the two is validated by design or testing. Cube specimens exhibit higher compressive strength in comparison with cylindrical specimens of UHPFRC due to the confinement effect, and the conversion factors for this difference have been obtained for UHPFRC by researchers [15][16]. Shaikh et al. [17] summarised the variation in compressive strength with the water/cement ratio for more than 70 different mixes of UHPFRC obtained from previous experimental works and thus it is not presented here. The compressive strengths ranged from 150 MPa to 300 MPa and the water/cement ratios were around 0.15–0.25. It was observed that as the water/cement ratio decreases, the compressive strength increases, which is generally true even for concrete with normal strength. 
The effects of specimen size on the mechanical properties of UHPFRC have to be considered when working with reduced scale specimens for laboratory testing and these effects must be considered in real structures. Thus, much research has been carried out to investigate the effect of the size of UHPFRC samples [18][19][20]. The results from these studies indicated that smaller samples possess higher compressive strengths and thus, the size effect of UHPFRC specimens is significant in the context of the compressive strength.
The compressive strength of UHPFRC is partly governed by the effect of pre-treatment. The rate of hydration in the mix can be enhanced by implementing proper heat treatment. The standard curing regime for UHPFRC includes steaming the specimens at 90 °C and 95% relative humidity for 2–6 days [21]. Thermally treated UHPFRC specimens possess a higher 28-day compressive strength compared to those of air-treated specimens [22]. In addition to heat treatment, the application of a confining pressure during the setting of UHPFRC can increase the compactness and thereby positively influence the compressive strength. Nevertheless, when considering the retrofitting applications of the existing structures, the curing regimes for additional strength gain are not pertinent.
The presence of steel fibres enhances the ductility and the tensile strength of UHPFRC. However, previous research has shown the addition of high amounts of steel fibres does not significantly influence the compressive strength enhancement of ultra-high performance concrete (UHPC) [23][24]. El dieb [25] observed that with the increase in the steel fibre volume fraction, the failure mode changed from sudden explosive failure to more characteristic of ductile failure, where the UHPFRC specimen was intact without chipping and spalling. Nevertheless, a significant improvement in compressive strength was not observed with the addition and increase in steel fibres. A slight increase in the compressive strength of UHPC was observed by Abbas et al. [26] with steel fibre addition, while the fibre length had a minimal effect on compressive strength. Introduction of steel fibres to UHPC can result in less entrapped air, increasing the density of the mix and resulting in increased compressive strength. In addition, a slight improvement in the compressive strength of UHPC could be due to the enhanced tolerance of lateral strains with the addition of steel fibres [27][28]. A negative impact of increasing the steel fibre concentration is fibre bundling, which can lead to weak spots, reducing the efficiency of fibres and the homogeneity of the mix. The effect of steel fibre shape and content on the mechanical properties of UHPFRC was investigated by Wu et al. [29]. Three different shapes of steel fibres were used in this study, namely straight, corrugated and hooked ends, and the fibre content was also varied by volume, ranging from 0 to 3%. Similar to the previous findings, the increase in the fibre content resulted in a minor increase in the compressive strength. In contrast, the shape of the steel fibre had a substantial effect on the compressive strength. For UHPFRC specimens with 3% hooked end and corrugated fibres, the increase in 28-day compressive strength was 48% and 59%, respectively, in comparison with specimens with the same amount of straight fibre. Thus, the effect of the steel fibre shape is paramount for compressive strength enhancement. 
Considering the economic and sustainable aspects, researchers have explored the inclusion of industrial by-products and even waste materials into UHPFRC matrices, without compromising the superior mechanical properties [30][31][32][33]. Granulated blast furnace slag, silica fume and fly ash have been used as partial clinker replacements. The results of the aforementioned experimental studies showed that UHPC mixes with lower volume fractions of partial clinker replacements had similar compressive strengths to the reference mix, with only with cement as the binder. Even the use of recycled glass cullets [34], waste ceramics [32] and waste bottom ash [33] did not result in significant reductions in the compressive strength when these materials are used in lower volume fractions in the UHPFRC matrix. One of the main drawbacks of UHPFRC is its high Portland cement content which increases the cost and also results in increased emission of greenhouse gases. Aldahdooh et al. [35] explored the possibility of adjusting the binder content in UHPFRC using the response surface method. It was found that for given water/binder and superplasticizer/cement ratios, the compressive strength did not rely on the binder content. Moreover, the capillary porosity increases with the increase in Portland cement and thus there is no strength enhancement with the increase in binder content.

2.2. Tensile and Flexural Strengths

The tensile and the flexural strengths of UHPFRC are significantly higher compared to conventional reinforced concrete. In UHPFRC, the improved tensile and flexural strengths are attributed to the dense particle packing and the addition of steel fibres. Shaikh et al. [17] presented a detailed summary of the tensile and flexural strength test results found in the literature for UHPFRC and thus they are not reviewed herein. Tensile strength was evaluated mostly using “dog-bone” specimens subjected to direct tension [36][37].
Researchers have explored the effect of varying the steel fibre content on the direct tensile and flexural tensile strengths of UHPFRC. It was found that the cracking and peak flexural tensile strength, as well as the strain-hardening behaviour, are improved by the addition of steel fibres compared to a UHPC mix without fibres [38]. Kang et al. [39] conducted notched three-point bending tests to investigate the flexural tensile strength of UHPFRC. A linear increase in the flexural tensile strength was observed when the fibre volume ratio increased from 0 to 5%. However, it was observed that the increase in the fibre content had little effect on the first crack strength and first crack deflection. The failed beam specimens exhibited a single vertical macro-crack along with multiple micro-cracks due to the effect of steel fibres [15]. The role of the fibres is to mitigate the spread of micro-cracks by absorbing the tensile stresses. Eiden et al. [40] observed that the splitting tensile strength increased from 34% to 67% for the steel fibre contents of 1% and 3%, respectively, compared to a UHPC mix without fibres. Additionally, a similar trend was found for the flexural strength. The flexural strengths improved from 15% to 40% with the increase in fibre content from 1% to 3%, respectively, in comparison with a mix without fibres. Wu et al. [29] varied the fibre content and the type of steel fibres in a UHPFRC mix to investigate the effects on the mechanical strength. Introduction of 2% straight, hooked end and corrugated fibres improved the flexural load by 46.3%, 81.1% and 61.4%. Furthermore, the peak deflection improvement was found to be 76.7%, 153.3% and 123.3%, respectively.
Park et al. [36] investigated the effects of blending fibres on the tensile behaviour of UHPFRC. Four types of steel macro-fibres and one type of micro-fibre were considered in this study. It was observed that the addition of micro-fibres into the hybrid system enhanced both the strain hardening and multiple cracking behaviours. The hybrid fibre, with a twisted fibre as the macro-fibre, inserted into UHPFRC produced the best strain hardening behaviour with an ultimate tensile strength of 18.6 MPa, with a corresponding strain of 0.64%. The performance ranking considering the post-cracking strength, strain capacity and multiple cracking behaviour was in the order of long smooth fibres < hooked end fibres < twisted fibres. In addition to the fibre type and content, the fibre orientation also has a considerable influence on the tensile and flexural strengths [41][42].
The size effect needs to be considered when evaluating both the flexural and tensile strengths. Similar to the compressive strength, decreased specimen sizes exhibit increased flexural strengths [20]. Moreover, the ductility is also enhanced in smaller specimens, possibly due to the improved fibre orientation in smaller specimens. Nguyen et al. [43] observed that the average number of cracks, and even the crack spacing, decreased with the decrease in specimen size. Frettlöhr et al. [44] explored the size effect and observed that for axial tension, the elastic limit as well as maximum tensile strength drastically reduced when the specimen depth increased from 25 to 100 mm. Similarly, for bending tests, the increase in prism height resulted in a decreased flexure tensile strength. Thus, the effect of specimen size cannot be neglected when considering the direct tensile and flexural tensile strengths.
In structural retrofitting applications, additional tensile and flexural strength enhancement via high-temperature curing is not applicable since it limits the application for precast UHPFRC products. Therefore, normal temperature curing has to be adopted for in situ construction. Nevertheless, previous research has shown that high temperature steam curing can improve both the tensile and flexural strengths [22]. Yang et al. [34] explored the effects of a curing regime on the mechanical properties of UHPFRC. A comparison of the specimens cured at 20 °C and 90 °C showed that a higher flexural strength was found in specimens cured at 90 °C. A flexural strength reduction of 10% was observed for specimens cured at 20 °C along with about 15% reduction in fracture energy compared with that of specimens cured at 90 °C.

2.3. Impact Strength

Impact strength is another aspect to be considered in structural retrofitting given the possibility of structures being subject to extreme loadings such as earthquakes, gas explosions and vehicle impact. Furthermore, in the recent past, blasts due to terrorist attacks also impart extreme loading on structures. When subjected to impacts, concrete undergoes elevated localised strain rates. Habel and Gauvreau [45] explored the rate-dependent behaviour of UHPFRC by conducting drop weight tests to apply dynamic three-point bending loading on UHPFRC plates. It was observed that both drop weight and quasi-static bending tests had identical failure modes. Ultimately, fracture occurred by fibre pullout in the centre of the specimen and a high moment region exhibited multiple cracking. Yoo et al. [46] examined the flexural behaviour of UHPFRC beams under low velocity impact loading. The experimental programme consisted of testing four large-sized beams using a drop-weight impact test machine. An improvement in the performances under impact loading was observed with the increase in the reinforcement ratio. With the increase in the reinforcement ratio, the maximum and the residual deflections of the beam decreased after the first impact. Thus, higher reinforcement ratios in beams exhibited a better performance for impact loading considering the deflections, crack widths and deflection recovery.

References

  1. Eide, M.B.; Hisdal, J.M. Ultra High Performance Fibre Reinforced Concrete (UHPFRC)–State of the Art; Blindern: Oslo, Norway, 2012.
  2. Richard, P.; Cheyrezy, M. Composition of Reactive Powder Concretes. Cem. Concr. Res. 1995, 25, 1501–1511.
  3. Rossi, P.; Arca, A.; Parant, E.; Fakhri, P. Bending and Compressive Behaviours of a New Cement Composite. Cem. Concr. Res. 2005, 35, 27–33.
  4. Tayeh, B.A.; Abu Bakar, B.H.; Megat Johari, M.A.; Voo, Y.L. Evaluation of Bond Strength between Normal Concrete Substrate and Ultra High Performance Fiber Concrete as a Repair Material. Proc. Procedia Eng. 2013, 54, 554–563.
  5. Rossi, P. Ultra-High Performance Fiber-Reinforced Concretes. Concr. Int. 2001, 23, 46–52.
  6. Habel, K. Structural Behaviour of Elements Combining Ultra-High Performance Fibre Reinforced Concretes (Uhpfrc) and Reinforced Concrete; EPEL: Lausanne, Switzerland, 2004; Volume 3036.
  7. Deeb, R.; Ghanbari, A.; Karihaloo, B.L. Development of Self-Compacting High and Ultra High Performance Concretes with and without Steel Fibres. Cem. Concr. Compos. 2012, 34, 185–190.
  8. Al-Osta, M.A. Exploitation of Ultrahigh-Performance Fibre-Reinforced Concrete for the Strengthening of Concrete Structural Members. Adv. Civ. Eng. 2018, 2018, 1–12.
  9. Abbas, S.; Nehdi, M.L.; Saleem, M.A. Ultra-High Performance Concrete: Mechanical Performance, Durability, Sustainability and Implementation Challenges. Int. J. Concr. Struct. Mater. 2016, 10, 271–295.
  10. de Larrard, F.; Sedran, T. Mixture-Proportioning of High-Performance Concrete. Cem. Concr. Res. 2002, 32, 1699–1704.
  11. Yu, R.; Spiesz, P.; Brouwers, H.J.H. Mix Design and Properties Assessment of Ultra-High Performance Fibre Reinforced Concrete (UHPFRC). Cem. Concr. Res. 2014, 56, 29–39.
  12. Ge, Z.; Wang, K.; Sun, R.; Huang, D.; Hu, Y. Properties of Self-Consolidating Concrete Containing Nano-CaCO3. J. Sustain. Cem. Based Mater. 2014, 3, 191–200.
  13. vande Voort, T.L. Design and Field Testing of Tapered H-Shaped Ultra High Performance Concrete Piles; Iowa State University ProQuest Dissertations Publishing: Iowa, IA, USA, 2008.
  14. Camacho, E.; López, J.Á.; Ros, P.S. Definition of Three Levels of Performance for UHPFRC-VHPFRC with Available Materials. In Proceedings of the Hipermat 2012—3rd International Symposium on UHPC and Nanotechnology for High Performance Construction Materials, Kassel, German, 7–12 March 2012.
  15. Graybeal, B.; Davis, M. Cylinder or Cube: Strength Testing of 80 to 200 MPa (11.6 to 29 Ksi) Ultra-High-Performance Fiber-Reinforced Concrete. ACI Mater. J. 2008, 105, 603.
  16. Kazemi, S.; Kazemi, S. Influence of Specimen Size and Fiber Content on Mechanical Properties of Ultra-High-Performance Fiber-Reinforced Concrete. ACI Mater. J. 2012, 109.
  17. Shaikh, F.U.A.; Luhar, S.; Arel, H.Ş.; Luhar, I. Performance Evaluation of Ultrahigh Performance Fibre Reinforced Concrete—A Review. Constr. Build. Mater. 2020, 232, 117152.
  18. Spasojevic, A.; Redaelli, D.; Fernández Ruiz, M.; Muttoni, A. Influence of Tensile Properties of UHPFRC on Size Effect in Bending. In Proceedings of the Second International Symposium on Ultra High Performance Concrete, Kassel, German, 5–7 March 2008.
  19. Wille, K.; Parra-Montesinos, G.J. Effect of Beam Size, Casting Method, and Support Conditions on Flexural Behavior of Ultra-High-Performance Fiber-Reinforced Concrete. ACI Mater. J. 2012, 109.
  20. Mahmud, G.H.; Yang, Z.; Hassan, A.M.T. Experimental and Numerical Studies of Size Effects of Ultra High Performance Steel Fibre Reinforced Concrete (UHPFRC) Beams. Constr. Build. Mater. 2013, 48, 1027–1034.
  21. Graybeal, B.A. Material Property Characterization of Ultra-High Performance Concrete; FHWA-HRT-06-103; U.S. Federal Highway Administration. Office of Infrastructure Research and Development: Washington, DC, USA, 2006.
  22. Ahlborn, T.M.; Misson, D.L.; Peuse, E.J.; Gilbertson, C.G.; Fehling, E. Durability and Strength Characterization of Ultra-High Performance Concrete under Variable Curing Regimes. In Proceedings of the 2nd International Symposium on Ultra High Performance Concrete, Kassel, Germany, 5–7 March 2008.
  23. Reda, M.M.; Shrive, N.G.; Gillott, J.E. Microstructural Investigation of Innovative UHPC. Cem. Concr. Res. 1999, 29, 323–329.
  24. Schmidt, M.; Fehling, E.; Teichmann, T.; Kai, B.; Roland, B. Ultra-High Performance Concrete: Perspective for the Precast Concrete Industry. Betonw. Und Fert.-Tech./Concr. Precast. Plant. Technol. 2003, 69, 16–29.
  25. El-Dieb, A.S. Mechanical, Durability and Microstructural Characteristics of Ultra-High-Strength Self-Compacting Concrete Incorporating Steel Fibers. Mater. Des. 2009, 30, 4286–4292.
  26. Abbas, S.; Soliman, A.M.; Nehdi, M.L. Exploring Mechanical and Durability Properties of Ultra-High Performance Concrete Incorporating Various Steel Fiber Lengths and Dosages. Constr. Build. Mater. 2015, 75, 429–441.
  27. Hassan, A.M.T.; Jones, S.W.; Mahmud, G.H. Experimental Test Methods to Determine the Uniaxial Tensile and Compressive Behaviour of Ultra High Performance Fibre Reinforced Concrete(UHPFRC). Constr. Build. Mater. 2012, 37, 874–882.
  28. Magureanu, C.; Sosa, I.; Negrutiu, C.; Heghes, B. Mechanical Properties and Durability of Ultra-High-Performance Concrete. ACI Mater. J. 2012, 109, 177–183.
  29. Wu, Z.; Shi, C.; He, W.; Wu, L. Effects of Steel Fiber Content and Shape on Mechanical Properties of Ultra High Performance Concrete. Constr. Build. Mater. 2016, 103, 8–14.
  30. van Tuan, N.; Ye, G.; van Breugel, K.; Fraaij, A.L.A.; Bui, D.D. The Study of Using Rice Husk Ash to Produce Ultra High Performance Concrete. Constr. Build. Mater. 2011, 25, 2030–2035.
  31. van Tuan, N.; Ye, G.; van Breugel, K.; Copuroglu, O. Hydration and Microstructure of Ultra High Performance Concrete Incorporating Rice Husk Ash. Cem. Concr. Res. 2011, 41, 1104–1111.
  32. Vejmelková, E.; Keppert, M.; Rovnaníková, P.; Ondráček, M.; Keršner, Z.; Černý, R. Properties of High Performance Concrete Containing Fine-Ground Ceramics as Supplementary Cementitious Material. Cem. Concr. Compos. 2012, 34, 55–61.
  33. Yu, R.; Spiesz, P.; Brouwers, H.J.H. Development of an Eco-Friendly Ultra-High Performance Concrete (UHPC) with Efficient Cement and Mineral Admixtures Uses. Cem. Concr. Compos. 2015, 55, 383–394.
  34. Yang, S.L.; Millard, S.G.; Soutsos, M.N.; Barnett, S.J.; Le, T.T. Influence of Aggregate and Curing Regime on the Mechanical Properties of Ultra-High Performance Fibre Reinforced Concrete (UHPFRC). Constr. Build. Mater. 2009, 23, 2291–2298.
  35. Aldahdooh, M.A.A.; Muhamad Bunnori, N.; Megat Johari, M.A. Evaluation of Ultra-High-Performance-Fiber Reinforced Concrete Binder Content Using the Response Surface Method. Mater. Des. 2013, 52, 957–965.
  36. Park, S.H.; Kim, D.J.; Ryu, G.S.; Koh, K.T. Tensile Behavior of Ultra High Performance Hybrid Fiber Reinforced Concrete. Cem. Concr. Compos. 2012, 34, 172–184.
  37. Wille, K.; El-Tawil, S.; Naaman, A.E. Properties of Strain Hardening Ultra High Performance Fiber Reinforced Concrete (UHP-FRC) under Direct Tensile Loading. Cem. Concr. Compos. 2014, 48, 53–66.
  38. Yoo, D.Y.; Lee, J.H.; Yoon, Y.S. Effect of Fiber Content on Mechanical and Fracture Properties of Ultra High Performance Fiber Reinforced Cementitious Composites. Compos. Struct. 2013, 106, 742–753.
  39. Kang, S.T.; Lee, Y.; Park, Y.D.; Kim, J.K. Tensile Fracture Properties of an Ultra High Performance Fiber Reinforced Concrete (UHPFRC) with Steel Fiber. Compos. Struct. 2010, 92, 61–71.
  40. Eldin, H.K.S.; Mohamed, H.A.; Khater, M.; Ahmed, S. Mechanical Properties of Ultra-High Performance Fiber Reinforced Concrete. Int. J. Eng. Innov. Technol. 2014, 4.
  41. Kang, S.T.; Lee, B.Y.; Kim, J.K.; Kim, Y.Y. The Effect of Fibre Distribution Characteristics on the Flexural Strength of Steel Fibre-Reinforced Ultra High Strength Concrete. Constr. Build. Mater. 2011, 25, 2450–2457.
  42. Yoo, D.Y.; Kang, S.T.; Yoon, Y.S. Effect of Fiber Length and Placement Method on Flexural Behavior, Tension-Softening Curve, and Fiber Distribution Characteristics of UHPFRC. Constr. Build. Mater. 2014, 64, 67–81.
  43. Nguyen, D.L.; Kim, D.J.; Ryu, G.S.; Koh, K.T. Size Effect on Flexural Behavior of Ultra-High-Performance Hybrid Fiber-Reinforced Concrete. Compos. B Eng. 2013, 45, 1104–1116.
  44. Frettlöhr, B.; Reineck, K.H.; Reinhardt, H.W. Size and Shape Effect of Uhpfrc Prisms Tested under Axial Tension and Bending. RILEM Bookseries 2012, 6, 365–372.
  45. Habel, K.; Gauvreau, P. Response of Ultra-High Performance Fiber Reinforced Concrete (UHPFRC) to Impact and Static Loading. Cem. Concr. Compos. 2008, 30, 938–946.
  46. Yoo, D.Y.; Banthia, N.; Kim, S.W.; Yoon, Y.S. Response of Ultra-High-Performance Fiber-Reinforced Concrete Beams with Continuous Steel Reinforcement Subjected to Low-Velocity Impact Loading. Compos. Struct. 2015, 126, 233–245.
More
Information
Subjects: Engineering, Civil
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 247
Revisions: 2 times (View History)
Update Date: 18 Apr 2023
1000/1000
Video Production Service