• Reference Manager
  • Simple TEXT file

People also looked at

Mini review article, new opportunities to advance the field of sports nutrition.

sports nutrition research paper

  • 1 Department of Physical Performance, Norwegian School of Sport Sciences, Oslo, Norway
  • 2 Gatorade Sports Science Institute, PepsiCo Life Sciences, Barrington, IL, United States
  • 3 Gatorade Sports Science Institute, PepsiCo Life Sciences, Global R&D, Leicestershire, United Kingdom
  • 4 Canadian Sport Institute-Pacific, Victoria, BC, Canada
  • 5 Exercise Science, Physical and Health Education, University of Victoria, Victoria, BC, Canada
  • 6 School of Sport and Health Sciences, University of Brighton, Eastbourne, United Kingdom

Sports nutrition is a relatively new discipline; with ~100 published papers/year in the 1990s to ~3,500+ papers/year today. Historically, sports nutrition research was primarily initiated by university-based exercise physiologists who developed new methodologies that could be impacted by nutrition interventions (e.g., carbohydrate/fat oxidation by whole body calorimetry and muscle glycogen by muscle biopsies). Application of these methods in seminal studies helped develop current sports nutrition guidelines as compiled in several expert consensus statements. Despite this wealth of knowledge, a limitation of the current evidence is the lack of appropriate intervention studies (e.g., randomized controlled clinical trials) in elite athlete populations that are ecologically valid (e.g., in real-life training and competition settings). Over the last decade, there has been an explosion of sports science technologies, methodologies, and innovations. Some of these recent advances are field-based, thus, providing the opportunity to accelerate the application of ecologically valid personalized sports nutrition interventions. Conversely, the acceleration of novel technologies and commercial solutions, especially in the field of biotechnology and software/app development, has far outstripped the scientific communities' ability to validate the effectiveness and utility of the vast majority of these new commercial technologies. This mini-review will highlight historical and present innovations with particular focus on technological innovations in sports nutrition that are expected to advance the field into the future. Indeed, the development and sharing of more “big data,” integrating field-based measurements, resulting in more ecologically valid evidence for efficacy and personalized prescriptions, are all future key opportunities to further advance the field of sports nutrition.

Introduction

Innovation has always been at the forefront of sport. Recent examples include drafting in cycling, clap skates in speed skating and more recently, carbon plate shoes in running. Sports nutrition is a relatively young discipline with <100 scientific papers published per year in the early 1990s, to about 3,500 per year today and a myriad of books ( Figure 1 ). Much of this progress was brought about by exercise physiologists who developed new methods and technologies within their laboratories (e.g., treadmills and ergometers) at universities around the world to study trained athletes (e.g., distance runners and cyclists) ( Hawley et al., 2015 ). Next to sports science, these developments promoted the emergence of another new discipline, that of sports nutrition. Some of the major innovations and corresponding knowledge milestones for sports nutrition research, combined with sports science research, are summarized in Figure 2 .

www.frontiersin.org

Figure 1 . Publications in Pubmed using the search term “Sports Nutrition” as of December, 2021 1 .

www.frontiersin.org

Figure 2 . Timeline of key innovations in Sports Nutrition and their respective influence on the field. *Including: genomics, transcriptomics, metabolomics, proteomics, phenomics, and other related omics (e.g., epigenomics).

In 2003, the International Olympic Committee (IOC) working group on sports nutrition concluded “The amount, composition, and timing of food intake can profoundly affect sports performance. Good nutritional practice will help athletes train hard, recover quickly and adapt more effectively with less risk of illness and injury” ( IOC Consensus Statement on Sports Nutrition, 2004 ). Nearing two decades later and these recommendations remain as pertinent. Despite all this progress, excitement and scientific endeavor, the ability to determine the impact of sports nutrition for different groups of athletes (e.g., different sports, ethnic groups, and sex) is still elusive. For example, there is substantial evidence for carbohydrate (CHO) ingestion before, during and after exercise ( Burke et al., 2011 ; Stellingwerff and Cox, 2014 ). However, it is difficult to separate the performance benefits of CHO ingestion per se vs. all other variables during competition (e.g., environment, competition, technology, equipment, and psychology).

Failure to establish the impact of any sports and exercise science intervention may result in a decline in recognition of the disciplines role in supporting the health and performance of athletes. This applies not only to the elite performers but also the young athlete, the exercising public, and the elderly. Therefore, this mini-review will focus on the opportunities to accelerate knowledge and practice of sports nutrition via the integration of technological innovations. We first highlight the present knowledge then propose ways of integrating technical advances and personalized prescription. Particular reference will be given to personalized prescriptions that are transforming and modernizing other life sciences.

The Present

Although, we primarily think of innovation in sports nutrition as directed at athlete performance outcomes, we also need to be innovative in our methods of synthesizing and dispersing research. The sports nutrition research data generated to date, have been compiled in several consensus documents on general sports ( Thomas D. T. et al., 2016 ), team sports ( Collins et al., 2021 ), and dietary supplements ( Maughan et al., 2018 ). However, a limitation of the current evidence base informing all sport science and sports medicine consensus guidelines is the lack of appropriate intervention studies (e.g., randomized controlled clinical trials) in elite populations ( McKay et al., 2022 ) that are ecologically valid (e.g., in real-life training and competition settings). Furthermore, often in practice, sports nutrition recommendations are retrospective fitted to sports in which the original research was not completed. For example much of the research to inform “stop-and-go” type sports has primarily been completed in soccer ( Williams and Rollo, 2015 ). There is also a primary focus on male subjects in the sports nutrition literature in general; yet most guidelines are “assumed” to be ideal for females as well. Furthermore, in the last 5 years, the focus has shifted from original research to reviews in the area of sports nutrition. Of the published articles, between 17.6 and 20.2% have been reviews (512–570), of which 4.4–6.0 meta-analyses (128–223). Some of this focus on reviews of all types rather than original research is enforced due to restrictions imposed due to Covid-19. However, there is a continuing need for original research in order to advance the discipline.

Consensus meetings and subsequent statements are fundamental in the generation of expert driven guidelines. Over the past 5 years, consensus statements in Sports Nutrition ranged between 0.3 and 0.5% ( Williams and Rollo, 2015 ; Jeukendrup, 2017 ; Pitsiladis et al., 2017 ; Sutehall et al., 2018 ; Burke et al., 2019 ; Stellingwerff et al., 2019 ; Baker et al., 2020 ; Muniz-Pardos et al., 2021 ) of the published articles. Historically, consensus statements, such as the International Olympic Committee (IOC) consensus on sports supplements ( Maughan et al., 2018 ) are drafted following in-person meetings of leading medical and scientific content experts. However, it is appreciated that such meetings can be costly in terms of budget, time and the carbon footprint of travel. More recently, major consensus statements have implemented remote online approaches, including the entire 2019 World Athletics (formerly IAAF) Nutrition for Athletics Consensus Update (featuring over 40 authors across 17 papers, e.g., 12). A remote consensus approach provides the opportunity to involve a wider contribution on topic guidelines rather than fewer selected opinion leaders. Establishing online working communication platforms as well as documents may also allow consensus documents to be updated with greater frequently or at pace with current literature. In summary, consensus statements informed by original research and meta-analyses, will require a greater reliance on new digital based approaches, while also respecting the need for in person meetings amongst experts.

Integrating Technical Advances to Field-Based Metrics

Most paradigms in sports nutrition have been established using laboratory experimentation, while neglecting evidence from in situ or field experimentation. This results in studies with limited ecological validity. In order to establish the efficacy of nutrition parameters for performance enhancement for all relevant populations, we need to better understand the competition demands of sport ( Stellingwerff et al., 2019 ). Recent advances in wearable technologies and real-time monitoring have accelerated the shift in research from the laboratory to the field in order to enhance ecological validity. This trend poses a real opportunity for all sport science disciplines, including sports nutrition, to embrace these technological developments. One such recent example was the implementation of live performance feedback of athletes (during 10,000 m, marathon, and race-walk events) competing in the heat at Tokyo 2020 ( Muniz-Pardos et al., 2021 ). Briefly, the aim of implementing this wireless technology during Tokyo Olympic Games was to help characterize the physiological and thermal strain experienced by athletes, as well as determine future management of athletes during a medical emergency as a result of a more timely and accurate diagnosis. The real-time monitoring comprised a smartwatch application, designed to collect, process and transmit a wide range of physiological, biomechanical, bioenergetics, and environmental data. This project was a success in terms of technological innovation but also general acceptance by athletes and sport's governing bodies. Such projects provide the opportunity for other new and valid sensors to assess performance- and health-related parameters particularly relevant to sports nutrition. One example is microfluidic technologies integrated into wearable patches to provide athletes instant feedback on sweat rate and sweat composition ( Baker et al., 2020 ). Wider adoption of such technologies will create more symbiotic relationships between sport, health and technology by harnessing the unique demands of elite sport (e.g., the need for unobtrusive devices that provide real-time feedback).

Given their symbiotic relationship, the evolution of sports nutrition, and sports science requires more holistic approaches with input from all major disciplines (e.g., coaching science, environmental physiology, and sports biomechanics), stakeholders, sponsors, and interested industry ( Pitsiladis et al., 2017 ). In recent years, physiology, nutrition, and technical advances have become increasingly integrated as part of new sport performance innovation strategies. A pertinent example is the Sub2 marathon project which was a novel proof-of-concept idea motivated by the need to focus on a holistic approach whilst promoting clean sport (i.e., high performance marathon running without doping) ( Pitsiladis et al., 2017 ). This was the first dedicated international research initiative launched in 2014 made up of multidisciplinary scientists from academia, elite athletes and strategic industry partners across many sport science and medicine domains. An exciting Sub2 innovation with particular sports nutrition focus was the carbohydrate “hydrogel” development. This innovative concept in sports drinks was tested in elite athletes training in Ethiopia and Kenya. The novel aspect of the gel was that it allowed runners to ingest and tolerate CHO concentrations much higher than would normally be possible to ingest while running (e.g., 30% CHO) ( Sutehall et al., 2018 ). This was important because a common challenge for runners is to meet CHO ingestion guidelines without experiencing gastrointestinal complaints ( Jeukendrup, 2017 ). This sports drink was subsequently trialed and tested in the field with positive response by elite runners during marathons and cyclists in the tour de France ( Sutehall et al., 2020 ). One laboratory-based study has confirmed improved running performance, greater carbohydrate oxidation and lower GI symptoms following hydrogel ingestion compared with a standard CHO solution ( Rowe et al., 2022 ). However, other laboratory-based studies have not reported any of these advantages following hydrogel ingestion compared to the ingestion of carbohydrate-electrolyte sport beverages ( Baur et al., 2019 ; King et al., 2020 ; McCubbin et al., 2020 ). Nevertheless, it is a great example of sports nutrition innovation specific to the needs of the sport in the field. The hydrogel innovation was adopted by both the breaking 2 , 3 . and INEOS 159 4 . projects to break the 2-h marathon barrier, a reflection of the perceived value of this putative innovation.

Combining emerging technologies are ideal to better our understanding of performance and to objectively test the impact of nutritional strategies in laboratory or real performance environments ( Table 1 ). Such innovations will also allow other sports, beyond the mainly studied endurance sports cycling and running, to be evaluated in terms of sports nutrition impact. The utilization of these technologies, and co-ordinated research, may allow for the rapid generation of large data sets across many other types of sport that have yet to be included in sports nutrition research. As such, this approach will (i) speed our knowledge of sports that are difficult to study, (ii) gain data from regional populations under-represented in the literature, and (iii) inform the advice of how specific nutrition guidelines maybe transferred to the field. Accordingly, Table 1 highlights examples of existing and emerging technologies and methodologies that are “field-based” and relatively non-invasive that may continue to drive and refine sports nutrition research, interventions and recommendations.

www.frontiersin.org

Table 1 . Examples of existing or potential “in field” non-invasive technologies or methodologies that may drive current or future nutrition studies, interventions, and/or recommendations.

The wearable/technological revolution promises in the near future to improve the ability to monitor a whole range of physiological parameters in the field. For example, apps, devices, and entire ecosystems are being developed and destined to improve the quality of dietary intake methods and therefore the accurately of athletes' daily energy intake (EI) ( Ferrara et al., 2019 ). These technological developments may enable the energy availability (EA) of individual athletes (i.e., calculated as EI–EEE/fat free mass) to be more accurately monitored. Correspondingly, a more comprehensively study of the athlete in situ would be possible. Thus, this approach represents an unprecedented opportunity to mitigate many unresolved issues in the field of sports nutrition such as relative energy deficiency in sport (RED-S) ( Mountjoy et al., 2018 ). Importantly, the recent explosion of wearable technology/apps/devices with often unsubstantiated claims require quality assurance standards for wearable devices. Such concerns have prompted the International Federation of Sports Medicine (FIMS) to create a global standard for wearable devices in sport and fitness ( Ash et al., 2020 , 2021 ). Organizations involved in sports nutrition also have the opportunity to engage in quality assurance processes to safeguard the credibility of the innovations in sports nutrition.

Personalized Prescriptions

There is no such thing as an “average” athlete. However, a key question is if there is an added value of personalized nutrition vs. general guidelines? Importantly, technology innovations will allow the individual response to a sports nutrition intervention to be determined. For instance, to find the individual recommendation of carbohydrate and fluid during exercise, we need knowledge about the energy demands of the sport, sweat losses, gastrointestinal limits, personal taste preferences and every element of the event. This needs to include research on different sport categories and target groups. This also presents the opportunity to follow athletes over a longer period of time, without associated human labor or time costs. For instance, to establish the extent to which an individual responds to different nutritional interventions, we need to conduct repeated testing in the same individual on several occasions. And the more complex the sport and its environment, the more test repetitions may be needed to establish the magnitude of impact of an intervention. It is also imperative that we determine athlete compliance with prescribed nutritional interventions. Such data will allow the evaluation of education and behavior change strategies, which may also provide opportunity for personalization.

The research on personalized sports nutrition will undoubtedly be the focus in the near future due to the technological advances in genomics technologies such as genetic sequencing. For instance, is has been suggested that the impact of DNA sequencing will become on a par with that of the microscope ( Shendure et al., 2019 ). Sports nutrition and sports science are encouraged to use these powerful technologies and to keep up with rapid developments to increase the chances of finding the best solutions possible. Such technologies are routinely used in biomedical research and precision medicine applications, such as for cancer, stroke and Alzheimer's disease, thus, vital lessons can be learned and transferred to sports nutrition. However, it is essential that these technological developments are not “oversold” and that their application in the field is founded on evidence-based research and not driven by commercial interests. At present, the use of genetic testing in both sports nutrition and sports science is at a very early stage. The consensus in the scientific literature being that genetic testing in sport science has very low clinical utility and should not be sold ( Guasch-Ferré et al., 2018 ; Tanisawa et al., 2020 ). This is in contrast to the ever-increasing number of companies selling genetic testing, supported by unfounded claims ( Webborn et al., 2015 ; Vlahovich et al., 2017 ; Tanisawa et al., 2020 ). The market value of genetic testing; USD 10.80 Billion in 2020, is forecast to reach USD 23.14 Billion by 2027 5 .

A more precision-based sports nutrition will also need to consider the other components of the “omics” cascade in addition to genomics (e.g., transcriptomics, metabolomics, proteomics, and single cell sequencing). Furthermore, such approaches may utilize powerful bioinformatics methods, such as machine learning and artificial intelligence to integrate the different layers of biological data required for understanding the functional consequences with the real time assessment of the “phenome” using 5G and 6G, sensors, devices and applications ( Mancin et al., 2021 ). The identification of relevant non-invasive biomarkers are attractive to athletes and practitioners, due to the speed and increased frequency of collection vs. traditional blood draws or questionnaires. However, these technologies should be adopted in accordance with ethical principles and within national/international regulatory frameworks, which require further development.

New Approaches to Fulfill Knowledge Gaps

Given recent technological breakthroughs, there are exciting opportunities for sports nutrition research to take gigantic leaps in the near future. Until now, most sports nutrition and sports physiology studies are performed in controlled laboratory environments and often study the effect of single nutrients. There is opportunity for sports nutrition research to embrace real world settings using real solutions and a more holistic approaches, such as performance benefits of whole foods, whole-body effects of low EA and “targeted nutritional periodization.” One example is a study using tracer technologies to compare the effect of whole eggs vs. egg whites on post exercise muscle protein synthesis ( van Vliet et al., 2017 ). New study designs should focus on real life settings that are strictly monitored with use of new technological advances, apps, and systems. As such, with a clear overview of nutritional demands of the sport and individual factors of impact, the extent of real-life effects of sports nutrition elements can be established.

Beyond the physiological impact of nutrients, there is also opportunity for sports nutrition research to study of cognitive and mental performance ( Habay et al., 2021 ). This shift will require sports nutrition researchers and nutritionists to adopt and further develop technological methods to allow the psychobiological determinants of performance to better defined. New research paradigms and technologies could revolutionize sports nutrition research from small landmark studies of the 1960s with mainly the authors as subjects taking muscle biopsies on themselves ( Bergström and Hultman, 1966 ), to the use of big data and collaboration between large groups of researchers. Examples of the latter are studies identifying genes implicated in hand grip strength involving over 195,000 subjects ( Willems et al., 2017 ) or investigating the effects of age, body composition, and sex on total expenditure by the doubly labeled water method in 6,421 participants from 29 countries ( Pontzer et al., 2021 ). The field of sports nutrition has the opportunity to adopt such collaborative practices combined with the application of the new and established technologies (see Table 1 ). It is reasonable to suggest that this approach will inevitably become the mainstay of personalized medicine, where treating the individual will be the norm rather than the average. If sports nutrition can embrace these challenges, it will thrive as an essential discipline and its relevance recognized in other fields ( Oikawa et al., 2021 ).

Limitations/Perspectives

While innovations are necessary and appealing, there needs to be a considered approach to implementation. Soon almost any parameter will be able to be measured or inferred, yet the use of such data especially during live performances remains to be explored. There also seems to be a trend toward 24/7 observations (e.g., Apple watch, Oura ring, WHOOP, and Biostrap). Caution is encouraged when moving from too little or no assessment to over monitored and scheduled, as a result of too much feedback and reliance on devices. For instance, the athlete should be focussing on racing/competition, not on heart rate or temperature or non-validated feedback directly from a device. Tracking may also be potentially stressful ( Andersen et al., 2020 ), albeit this remains to be determined in athlete populations. When evolving sports nutrition research with new technological advances, it is important to continuously question the application to practice as well as the reliability and reliance of devices.

The integration of new technologies in elite populations will also require closer collaborations between research and practitioners, and then directly to the athlete and coach ( Bartlett and Drust, 2021 ). However, multidisciplinary sport science and medicine teams do not come without challenges and clear communication, roles and responsibilities are essential ( Dijkstra et al., 2014 ) with the athlete and coach at the centre of accountability.

Finally, impactful implementation of these innovations and technological developments especially in elite athletic populations is going to require the continued and better integration of behavioral change psychology in sports nutrition. A recent systematic review highlighted some of the most effective behavioral strategies used in sports nutrition ( Bentley et al., 2020 ).

Conclusions

Innovation is at the core of sports nutrition research and has pushed the field forward even before sports nutrition was recognized as a separate discipline. We are at a critical juncture in the evolution of this discipline primed to utilize new technologies to support the success of specific sports and individual athletes. Sharing data in new and more efficient ways, integrating field based physiological measures, and personalized prescriptions are key opportunities to advance sports nutrition. However, technological advances should not be used in haste and must first be evaluated to determine their functionality and value to the athletes health and performance. In summary, nutrition is but one of many complex and integrated sport performance determinants, and the impact of any new intervention should be assessed along a risk-reward continuum.

Author Contributions

All authors listed have made a substantial, direct, and intellectual contribution to the work and approved it for publication.

Author Disclaimer

The views expressed in this article are those of the authors and do not necessarily reflect the position or policy of PepsiCo, Inc.

Conflict of Interest

IR and MK are employees of the Gatorade Sports Science Institute, a division of PepsiCo, Inc. KJ, TS, and YP, received speaking honoraria, for the GSSI ECSS 2021 pre-conference symposium which inspired this article. YP is the founding member of the Sub2 project ( www.sub2hrs.com ). The Sub2 project is affiliated to a non-trading company (Athlome Limited, UK) that is minor (<1.1%) shareholder of Maurten AB, Gothenburg, Sweden.

Publisher's Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

1. ^ https://pubmed.ncbi.nlm.nih.gov/?term=sports+nutrition&filter=years.1946-2021&timeline=expanded&sort=date&sort_order=asc .

2. ^ www.runnersworld.com/uk/news/a29100149/breaking2 .

3. ^ www.nike.com/gb/running/breaking2 .

4. ^ http://www.ineos159challenge.com/ .

5. ^ https://brandessenceresearch.com/requestSample/PostId/1362 .

Andersen, T. O., Langstrup, H., and Lomborg, S. (2020). Experiences with wearable activity data during self-care by chronic heart patients: qualitative study. J. Med. Internet Res. 22:e15873. doi: 10.2196/15873

PubMed Abstract | CrossRef Full Text | Google Scholar

Ash, G. I., Stults-Kolehmainen, M., Busa, M. A., Gaffey, A. E., Angeloudis, K., Muniz-Pardos, B., et al. (2021). Establishing a global standard for wearable devices in sport and exercise medicine: perspectives from academic and industry stakeholders. Sports Med . 51, 2237–2250. doi: 10.1007/s40279-021-01543-5

Ash, G. I., Stults-Kolehmainen, M., Busa, M. A., Gregory, R., Garber, C. E., Liu, J., et al. (2020). Establishing a global standard for wearable devices in sport and fitness: perspectives from the New England Chapter of the American College of Sports Medicine Members. Curr. Sports Med. Rep . 19, 45–49. doi: 10.1249/JSR.0000000000000680

Baker, L. B., Model, J. B., Barnes, K. A., Anderson, M. L., Lee, S. P., Lee, K. A., et al. (2020). Skin-interfaced microfluidic system with personalized sweating rate and sweat chloride analytics for sports science applications. Sci. Adv . 6, eabe3929. doi: 10.1126/sciadv.abe3929

Bartlett, J. D., and Drust, B. (2021). A framework for effective knowledge translation and performance delivery of Sport Scientists in professional sport. Eur. J. Sport Sci . 21, 1579–1587. doi: 10.1080/17461391.2020.1842511

Baur, D. A., Toney, H. R., Saunders, M. J., Baur, K. G., Luden, N. D., and Womack, C. J. (2019). Carbohydrate hydrogel beverage provides no additional cycling performance benefit versus carbohydrate alone. Eur. J. Appl. Physiol . 119, 2599–2608. doi: 10.1007/s00421-019-04240-4

Bentley, M. R. N., Mitchell, N., and Backhouse, S. H. (2020). Sports nutrition interventions: a systematic review of behavioural strategies used to promote dietary behaviour change in athletes. Appetite 150, 104645. doi: 10.1016/j.appet.2020.104645

Bergström, J., and Hultman, E. (1966). Muscle glycogen synthesis after exercise: an enhancing factor localized to the muscle cells in man. Nature 210, 309–310. doi: 10.1038/210309a0

Burke, L. M., Castell, L. M., Casa, D. J., Close, G. L., Costa, R. J. S., Desbrow, B., et al. (2019). International Association of Athletics Federations Consensus Statement 2019: nutrition for athletics. Int. J. Sport Nutr. Exerc. Metab . 29, 73–84. doi: 10.1123/ijsnem.2019-0065

Burke, L. M., Hawley, J. A., Wong, S. H. S., and Jeukendrup, A. E. (2011). Carbohydrates for training and competition. J. Sports Sci . 29(Suppl. 1), 17–S27. doi: 10.1080/02640414.2011.585473

Collins, J., Maughan, R. J., Gleeson, M., Bilsborough, J., Jeukendrup, A., Morton, J. P., et al. (2021). UEFA expert group statement on nutrition in elite football. Current evidence to inform practical recommendations and guide future research. Br. J. Sports Med . 55, 416. doi: 10.1136/bjsports-2019-101961

Décombaz, J., Beaumont, M., Vuichoud, J., Bouisset, F., and Stellingwerff, T. (2012). Effect of slow-release β-alanine tablets on absorption kinetics and paresthesia. Amino Acids 43, 67–76. doi: 10.1007/s00726-011-1169-7

Dijkstra, H. P., Pollock, N., Chakraverty, R., and Alonso, J. M. (2014). Managing the health of the elite athlete: a new integrated performance health management and coaching model. Br. J. Sports Med . 48, 523–531. doi: 10.1136/bjsports-2013-093222

Duggleby, S. L., and Waterlow, J. C. (2005). The end-product method of measuring whole-body protein turnover: a review of published results and a comparison with those obtained by leucine infusion. Br. J. Nutr . 94, 141–153. doi: 10.1079/BJN20051460

Ferrara, G., Kim, J., Lin, S., Hua, J., and Seto, E. (2019). A focused review of smartphone diet-tracking apps: usability, functionality, coherence with behavior change theory, and comparative validity of nutrient intake and energy estimates. JMIR Mhealth Uhealth 7, e9232. doi: 10.2196/mhealth.9232

Goffinet, L., Barrea, T., Beauloye, V., and Lysy, P. A. (2017). Blood versus urine ketone monitoring in a pediatric cohort of patients with type 1 diabetes: a crossover study. Ther. Adv. Endocrinol. Metab . 8, 3–13. doi: 10.1177/2042018816681706

Guasch-Ferré, M, Dashti, H. S., and Merino, J. (2018). Nutritional genomics and direct-to-consumer genetic testing: an overview. Adv. Nutr . 9, 128–135. doi: 10.1093/advances/nmy001

Haakonssen, E. C., Martin, D. T., Burke, L. M., and Jenkins, D. G. (2013). Energy expenditure of constant- and variable-intensity cycling: power meter estimates. Med. Sci. Sports Exerc . 45, 1833–1840. doi: 10.1249/MSS.0b013e31828e18e6

Habay, J., Van Cutsem, J., Verschueren, J., De Bock, S., Proost, M., De Wachter, J., et al. (2021). Mental fatigue and sport-specific psychomotor performance: a systematic review. Sports Med . 51, 1527–1548. doi: 10.1007/s40279-021-01429-6

Hawley, J. A., Maughan, R. J., and Hargreaves, M. (2015). Exercise metabolism: historical perspective. Cell Metab . 22, 12–17. doi: 10.1016/j.cmet.2015.06.016

IOC Consensus Statement on Sports Nutrition (2004). Available online at: https://stillmed.olympic.org/Documents/Commissions_PDFfiles/Medical_commission/IOC_CONSENSUS_STATEMENT_ON_SPORTS_NUTRITION_2003.pdf (accessed December 14, 2021).

Jeukendrup, A. E.. (2010). Carbohydrate and exercise performance: the role of multiple transportable carbohydrates. Curr. Opin. Clin. Nutr. Metab. Care 13, 452–457. doi: 10.1097/MCO.0b013e328339de9f

Jeukendrup, A. E.. (2017). Training the gut for athletes. Sports Med. 47(Suppl. 1), 101–110. doi: 10.1007/s40279-017-0690-6

PubMed Abstract | CrossRef Full Text

King, A. J., Rowe, J. T., and Burke, L. M. (2020). Carbohydrate hydrogel products do not improve performance or gastrointestinal distress during moderate-intensity endurance exercise. Int. J. Sport Nutr. Exerc. Metab . 30, 305–314. doi: 10.1123/ijsnem.2020-0102

Lee, E. C., Fragala, M. S., Kavouras, S. A., Queen, R. M., Pryor, J. L., and Casa, D. J. (2017). Biomarkers in sports and exercise: tracking health, performance, and recovery in athletes. J. Strength Cond. Res . 31, 2920–2937. doi: 10.1519/JSC.0000000000002122

Mancin, L., Rollo, I., Mota, J. F., Piccini, F., Carletti, M., Susto, G. A., et al. (2021). Optimizing microbiota profiles for athletes. Exerc. Sport Sci. Rev . 49, 42–49. doi: 10.1249/JES.0000000000000236

Maughan, R. J., Burke, L. M., Dvorak, J., Larson-Meyer, D. E., Peeling, P., Phillips, S. M., et al. (2018). IOC consensus statement: dietary supplements and the high-performance athlete. Br. J. Sports Med . 52, 439–455. doi: 10.1136/bjsports-2018-099027

McCubbin, A. J., Zhu, A., Gaskell, S. K., and Costa, R. J. S. (2020). Hydrogel carbohydrate-electrolyte beverage does not improve glucose availability, substrate oxidation, gastrointestinal symptoms or exercise performance, compared with a concentration and nutrient-matched placebo. Int. J. Sport Nutr. Exerc. Metab . 30, 25. doi: 10.1123/ijsnem.2019-0090

McKay, A. K. A., Stellingwerff, T., Smith, E. S., Martin, D. T., Mujika, I., Goosey-Tolfrey, V. L., et al. (2022). Defining training and performance calibre: a participant classification framework. Int. J. Sports Physiol. Perform . 1–15. doi: 10.1123/ijspp.2021-0451

Mountjoy, M., Sundgot-Borgen, J. K., Burke, L. M., Ackerman, K. E., Blauwet, C., Constantini, N., et al. (2018). IOC consensus statement on relative energy deficiency in sport (RED-S): 2018 update. Br. J. Sports Med . 52, 687–697. doi: 10.1136/bjsports-2018-099193

Muniz-Pardos, B., Angeloudis, K., Guppy, F. M., Keramitsoglou, I., Sutehall, S., Bosch, A., et al. (2021). Wearable and telemedicine innovations for Olympic events and elite sport. J. Sports Med. Phys. Fitness 61, 1061–1072. doi: 10.23736/S0022-4707.21.12752-5

Nieves, J. W., Melsop, K., Curtis, M., Kelsey, J. L., Bachrach, L. K., Greendale, G., et al. (2010). Nutritional factors that influence change in bone density and stress fracture risk among young female cross-country runners. PM R 2, 740–750; quiz 794. doi: 10.1016/j.pmrj.2010.04.020

O'Driscoll, R., Turicchi, J., Beaulieu, K., Scott, S., Matu, J., Deighton, K., et al. (2020). How well do activity monitors estimate energy expenditure? A systematic review and meta-analysis of the validity of current technologies. Br. J. Sports Med . 54, 332–340. doi: 10.1136/bjsports-2018-099643

Oikawa, S. Y., Brisbois, T. D., van Loon, L. J. C., and Rollo, I. (2021). Eat like an athlete: insights of sports nutrition science to support active aging in healthy older adults. Geroscience 43, 2485–2495. doi: 10.1007/s11357-021-00419-w

Pitsiladis, Y., Ferriani, I., Geistlinger, M., de Hon, O., Bosch, A., and Pigozzi, F. (2017). A holistic antidoping approach for a fairer future for sport. Curr. Sports Med. Rep . 16, 222–224. doi: 10.1249/JSR.0000000000000384

Pontzer, H., Yamada, Y., Sagayama, H., Ainslie, P. N., Andersen, L. F., Anderson, L. J., et al. (2021). Daily energy expenditure through the human life course. Science 373, 808–812. doi: 10.1126/science.abe5017

Reisz, J. A., and D'Alessandro, A. (2017). Measurement of metabolic fluxes using stable isotope tracers in whole animals and human patients. Curr. Opin. Clin. Nutr. Metab. Care 20, 366–374. doi: 10.1097/MCO.0000000000000393

Rowe, J. T., King, R. F. G. J., King, A. J., Morrison, D. J., Preston, T., Wilson, O. J., et al. (2022). Glucose and fructose hydrogel enhances running performance, exogenous carbohydrate oxidation, and gastrointestinal tolerance. Med. Sci. Sports Exerc . 54, 129–140. doi: 10.1249/MSS.0000000000002764

Shendure, J., Balasubramanian, S., Church, G. M., Gilbert, W., Rogers, J., Schloss, J. A., et al. (2019). Publisher correction: DNA sequencing at 40: past, present and future. Nature 568:E11. doi: 10.1038/s41586-019-1120-8

Speakman, J. R., and Hambly, C. (2016). Using doubly-labelled water to measure free-living energy expenditure: some old things to remember and some new things to consider. Comp. Biochem. Physiol. A Mol. Integr. Physiol . 202, 3–9. doi: 10.1016/j.cbpa.2016.03.017

Stellingwerff, T., and Cox, G. R. (2014). Systematic review: carbohydrate supplementation on exercise performance or capacity of varying durations. Appl. Physiol. Nutr. Metab . 39, 998–1011. doi: 10.1139/apnm-2014-0027

Stellingwerff, T., Morton, J. P., and Burke, L. M. (2019). A framework for periodized nutrition for athletics. Int. J. Sport Nutr. Exerc. Metab . 29, 141–151. doi: 10.1123/ijsnem.2018-0305

Surapongchai, J., Saengsirisuwan, V., Rollo, I., Rendell, R. K., Nithitsuttibuta, K., Sainiyom, P., et al. (2021). Hydration status, fluid intake, sweat rate, and sweat sodium concentration in recreational tropical native runners. Nutrients 13:1374. doi: 10.3390/nu13041374

Sutehall, S., Galloway, S. D. R., Bosch, A., and Pitsiladis, Y. (2020). Addition of an alginate hydrogel to a carbohydrate beverage enhances gastric emptying. Med. Sci. Sports Exerc . 8, 1785–1792. doi: 10.1249/MSS.0000000000002301

Sutehall, S., Muniz-Pardos, B., Bosch, A., Di Gianfrancesco, A., and Pitsiladis, Y. (2018). Sports drinks on the edge of a new era. Curr. Sports Med. Rep . 17, 112–116. doi: 10.1249/JSR.0000000000000475

Tanisawa, K., Wang, G., Seto, J., Verdouka, I., Twycross-Lewis, R., Karanikolou, A., et al. (2020). Sport and exercise genomics: the FIMS 2019 consensus statement update. Br. J. Sports Med . 54, 969–975. doi: 10.1136/bjsports-2019-101532

Thomas, D. T., Erdman, K. A., and Burke, L. M. (2016). American College of Sports Medicine Joint Position Statement. Nutrition and athletic performance. Med. Sci. Sports Exerc . 48, 543–568. doi: 10.1249/MSS.0000000000000852

Thomas, F., Pretty, C. G., Desaive, T., and Chase, J. G. (2016). Blood glucose levels of subelite athletes during 6 days of free living. J. Diabetes Sci. Technol . 10, 1335–1343. doi: 10.1177/1932296816648344

van Vliet, S., Shy, E. L., Sawan, S. A., Beals, J. W., West, D. W., Skinner, S. K., et al. (2017). Consumption of whole eggs promotes greater stimulation of postexercise muscle protein synthesis than consumption of isonitrogenous amounts of egg whites in young men. Am. J. Clin. Nutr . 106, 1401–1412. doi: 10.3945/ajcn.117.159855

Vlahovich, N., Hughes, D., Griffiths, L. R., Wang, G., Pitsiladis, Y. P., and Eynon, N. (2017). Genetic testing for exercise prescription and injury prevention: AIS-Athlome consortium joint statement. BMC Genomics 18(Suppl 8), 818. doi: 10.1186/s12864-017-4185-5

Webborn, N., Williams, A., McNamee, M., Bouchard, C., Pitsiladis, Y., Ahmetov, I., et al. (2015). Direct-to-consumer genetic testing for predicting sports performance and talent identification: Consensus statement. Br. J. Sports Med . 49, 1486–1491. doi: 10.1136/bjsports-2015-095343

Willems, S. M., Wright, D. J., Day, F. R., Trajanoska, K., Joshi, P. K., Morris, J. A., et al. (2017). Large-scale GWAS identifies multiple loci for hand grip strength providing biological insights into muscular fitness. Nat. Commun . 8, 16015. doi: 10.1038/ncomms16015

Williams, C., and Rollo, I. (2015). Carbohydrate nutrition and team sport performance. Sports Med . 45(Suppl 1), S13–S22. doi: 10.1007/s40279-015-0399-3

Keywords: innovation, wearables, technology, performance, health, diet, wellness, athletes

Citation: Jonvik KL, King M, Rollo I, Stellingwerff T and Pitsiladis Y (2022) New Opportunities to Advance the Field of Sports Nutrition. Front. Sports Act. Living 4:852230. doi: 10.3389/fspor.2022.852230

Received: 10 January 2022; Accepted: 18 January 2022; Published: 17 February 2022.

Reviewed by:

Copyright © 2022 Jonvik, King, Rollo, Stellingwerff and Pitsiladis. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Yannis Pitsiladis, y.pitsiladis@brighton.ac.uk

This article is part of the Research Topic

Insights in Sport and Exercise Nutrition: 2021

Sports Nutrition for Optimal Athletic Performance and Health: Old, New and Future Perspectives

  • Open access
  • Published: 06 November 2019
  • Volume 49 , pages 99–101, ( 2019 )

Cite this article

You have full access to this open access article

  • Lawrence L. Spriet 1  

12k Accesses

3 Citations

52 Altmetric

Explore all metrics

Avoid common mistakes on your manuscript.

This supplement examines several sports nutrition topics related to the optimisation of athletic performance and health. Some of the research areas have been examined for many years but are presently being examined with new methods and perspectives, while others are very new and some may be seen as approaches for the future. The authors have provided an excellent explanation of these new approaches and perspectives, as well as outlining where additional sports nutrition research is needed—especially with athletic populations and the long-term health of athletes. This collection of papers makes it very clear how important adequate nutrition is to the performance and well-being of athletes and how far reaching the negative effects of low energy availability can be, with these topics being discussed in five out of the eight papers in this supplement.

The Gatorade Sports Science Institute (GSSI) has been bringing sports nutrition and sports science researchers together for over 30 years to address and discuss many topics that relate to the health and performance of athletes. Since 2012, these meetings have been known as the GSSI Expert Panel. The latest meeting in March of 2019 was held to discuss old, new and future perspectives related to athlete nutritional, performance and health issues. Following the meeting, the authors summarized the recent work in their topic area, resulting in the manuscripts in this Sports Medicine supplement (the seventh in a series supported by GSSI).

The initial two papers highlight areas relevant to athletic performance that are difficult to study—does hypohydration really impair endurance performance [ 1 ] and what causes muscle cramping [ 2 ]? In the former situation, how does the experimenter effectively blind the research subject from knowing their hydration status? In addition, inducing hypohydration can be uncomfortable and unfamiliar to the subjects, with both problems potentially leading to performance decrements that are unrelated to hypohydration. The authors discuss recent attempts to rectify these problems using blinded hydration methods and conclude that hypohydration of ~ 2–3% body mass decreases endurance cycling performance in the heat, at least when no/little fluid is ingested [ 1 ]. In the muscle cramping situation, the authors have carefully described that water and salt balance are involved in cramping, using data from early studies in industrial settings with many subjects and also more recent work with smaller numbers of athletes [ 2 ]. In other situations, however, sustained abnormal spinal reflex activity seems to be the cause. Since no laboratory experimental models appear to be applicable to whole body exercise situations where muscle cramps occur, the authors argue that a single strategy for prevention or treatment will not be found.

In the next paper, De Souza et al. [ 3 ] provide a brief overview of the Female Athlete Triad and an update on the current thinking regarding energy availability. They also discuss the available literature relevant to a similar syndrome in males that is referred to as the Male Athlete Triad. To date, it appears that the energetic, reproductive and bone systems in men are more resilient to the effects of low energy availability compared to those of women, requiring more severe perturbations before alterations are observed. In addition, recovery of the hypothalamic pituitary gonadal axis occurs more quickly in men than in women. However, far more research with males experiencing low energy availability is needed.

The paper on nutrition and athlete bone health [ 4 ] also stresses the need for more athlete-specific research, especially as it relates to longer-term bone health (e.g., risk of osteopenia and osteoporosis) and shorter-term risk of bony injuries. Bone is a nutritionally modified tissue and generally benefits from weight-bearing activities, although not all athletes engage in weight-bearing sports. While nutritional requirements to support bone health may not be different between athletes and the general population, the authors highlight situations that may be relevant for athletes, including low energy availability, low carbohydrate availability, protein intake, vitamin D intake and dermal calcium and sodium losses. The paper by Walsh [ 5 ] outlines a new perspective on nutrition and athlete health to better understand how sick an athlete will become when they get an infection. This paradigm includes the concepts of immune resistance (destroying microbes) and immune tolerance (dampening defence but controlling infection to a non-damaging level). It also suggests that research efforts on nutritional supplements that may provide immunological tolerance and reduce the infection burden in athletes are needed.

The paper on nutrition and health at altitude [ 6 ] has been written by several scientists renowned for their work examining strategies to enhance adaptation, improve performance and maintain health in athletes living and training at low-to-moderate altitudes (1600–2400 m). Much of the existing altitude research was conducted at high to extreme altitudes (> 3000 m) and not the lower altitudes that athletes typically train at. While the authors highlight several nutritional issues that must be monitored at altitude, they stress that special attention must be given to the possibility of poor energy availability and increased iron requirements limiting the adaptations to altitude. Also, to deal with the possibility of increased oxidative stress at altitude, foods rich in antioxidants are recommended rather than high-dose antioxidant supplements.

The final two papers examine approaches to athlete nutrition and performance that might be called futuristic. The examination of blood test data, as a physiological profiling and monitoring tool, is becoming more routinely used in professional and elite high-performance athletes [ 7 ]. Much useful information can be obtained from blood tests, including the identification of iron, vitamin and energy deficiency, the identification of oxidative stress and inflammation status and the characteristics of red blood cell populations. Such data can be used to identify the effectiveness of training interventions, nutritional strategies and training load tolerance. The authors discuss perspectives, limitations and recommendations for sports science and sports medicine practitioners, who may use athlete blood profiling and monitoring for nutrition and performance purposes. In the final paper, Joyner [ 8 ] discusses the physiological determinants of human endurance performance, maximal oxygen uptake, the lactate threshold and running economy or effciency. He examines the genetics of endurance performance, as many of us may assume that individual differences in our genetic endowment would account for differences in endurance performance. However, he concludes that at present, interindividual differences in DNA sequence explain only a small fraction of the physiology underpinning sports performance.

The papers of this supplement have identified several areas of sports nutrition research that need to be studied or restudied to optimize sports performance and health for athletes. While we study groups of athletes, it is also clear that there are large individual variations between athletes and that we lack research in many areas for female athletes and in some cases male athletes. It is hoped that these papers have provided interesting perspectives on old, new and future areas of sports nutrition research and will spur additional research in these areas.

James LJ, Funnell MP, James RM, Mears SA. Does hypohydration really impair endurance performance? Methodological considerations for interpreting hydration research. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01188-5 (Suppl) .

Article   PubMed   Google Scholar  

Maughan RJ, Shirreffs SM. Muscle cramping during exercise: causes, solutions and questions remaining. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01162-1 (Suppl) .

De Souza MJ, Koltun KJ, Williams NI. The role of energy availability in reproductive function in the female athlete triad and extension of its effects to men: an initial working model of a similar syndrome in male athletes. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01217-3 (Suppl) .

Sale C, Elliott-Sale KJ. Nutrition and athlete bone health. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01161-2 (Suppl) .

Walsh NP. Nutrition and athlete immune health: new perspectives on an old paradigm. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01160-3 (Suppl) .

Stellingwerff T, Peeling P, Garvican-Lewis LA, Hall R, Koivisto AE, Heikura IA, Burke LM. Nutrition and altitude: strategies to enhance adaptation, improve performance and maintain health: a narrative review. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01159-w (Suppl) .

Pedlar CR, Newell J, Lewis NA. Blood biomarker profiling and monitoring for high performance physiology and nutrition: current perspectives, limitations and recommendations. Sports Med. 2019. https://doi.org/10.1007/s40279-019-01158-x (Suppl) .

Joyner MJ. Genetic approaches for sports performance: how far away are we? Sports Med. 2019. https://doi.org/10.1007/s40279-019-01164-z (Suppl) .

Download references

Acknowledgements

This supplement is supported by the Gatorade Sports Science Institute (GSSI). The supplement was guest edited by Dr. Lawrence L. Spriet, who attended a meeting of the GSSI Expert Panel in March 2019 and received honoraria from the GSSI, a division of PepsiCo, Inc., for his participation in the meeting and the writing of this preface. Dr. Spriet received no honorarium for guest editing the supplement. Dr. Spriet suggested peer reviewers for each paper, which were sent to the Sports Medicine Editor-in-Chief for approval, prior to any reviewers being approached. Dr. Spriet provided comments on each paper and made an editorial decision based on comments from the peer reviewers and the Editor-in-Chief. Where decisions were uncertain, Dr. Spriet consulted with the Editor-in-Chief.

Author information

Authors and affiliations.

Human Health and Nutritional Sciences, University of Guelph, Guelph, ON, N1G 2W1, Canada

Lawrence L. Spriet

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Lawrence L. Spriet .

Ethics declarations

An honorarium for preparation of this article was provided by the GSSI.

Conflict of Interest

Lawrence Spriet has no conflicts of interest relevant to the content of this article.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License ( http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Spriet, L.L. Sports Nutrition for Optimal Athletic Performance and Health: Old, New and Future Perspectives. Sports Med 49 (Suppl 2), 99–101 (2019). https://doi.org/10.1007/s40279-019-01224-4

Download citation

Published : 06 November 2019

Issue Date : December 2019

DOI : https://doi.org/10.1007/s40279-019-01224-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Advertisement

  • Find a journal
  • Publish with us
  • Track your research

From Paper to Podium: Quantifying the Translational Potential of Performance Nutrition Research

Affiliations.

  • 1 Research Institute for Sport and Exercise Science, Liverpool John Moores University, Liverpool, L3 3AF, UK. [email protected].
  • 2 Research Institute for Sport and Exercise Science, Liverpool John Moores University, Liverpool, L3 3AF, UK.
  • PMID: 30671902
  • PMCID: PMC6445818
  • DOI: 10.1007/s40279-018-1005-2

Sport nutrition is one of the fastest growing and evolving disciplines of sport and exercise science, demonstrated by a 4-fold increase in the number of research papers between 2012 and 2018. Indeed, the scope of contemporary nutrition-related research could range from discovery of novel nutrient-sensitive cell-signalling pathways to the assessment of the effects of sports drinks on exercise performance. For the sport nutrition practitioner, the goal is to translate innovations in research to develop and administer practical interventions that contribute to the delivery of winning performances. Accordingly, step one in the translation of research to practice should always be a well-structured critique of the translational potential of the existing scientific evidence. To this end, we present an operational framework (the "Paper-2-Podium Matrix") that provides a checklist of criteria for which to prompt the critical evaluation of performance nutrition-related research papers. In considering the (1) research context, (2) participant characteristics, (3) research design, (4) dietary and exercise controls, (5) validity and reliability of exercise performance tests, (6) data analytics, (7) feasibility of application, (8) risk/reward and (9) timing of the intervention, we aimed to provide a time-efficient framework to aid practitioners in their scientific appraisal of research. Ultimately, it is the combination of boldness of reform (i.e. innovations in research) and quality of execution (i.e. ease of administration of practical solutions) that is most likely to deliver the transition from paper to podium.

Publication types

  • Athletic Performance
  • Decision Making
  • Patient Selection
  • Research Design
  • Sports Nutritional Physiological Phenomena*
  • Translational Research, Biomedical*
  • Open access
  • Published: 02 February 2010

ISSN exercise & sport nutrition review: research & recommendations

  • Richard B Kreider 1 ,
  • Colin D Wilborn 2 ,
  • Lem Taylor 2 ,
  • Bill Campbell 3 ,
  • Anthony L Almada 4 ,
  • Rick Collins 5 ,
  • Mathew Cooke 6 ,
  • Conrad P Earnest 7 ,
  • Mike Greenwood 8 ,
  • Douglas S Kalman 9 ,
  • Chad M Kerksick 10 ,
  • Susan M Kleiner 11 ,
  • Brian Leutholtz 8 ,
  • Hector Lopez 12 ,
  • Lonnie M Lowery 13 ,
  • Ron Mendel 14 ,
  • Abbie Smith 10 ,
  • Marie Spano 15 ,
  • Robert Wildman 16 ,
  • Darryn S Willoughby 8 ,
  • Tim N Ziegenfuss 17 &
  • Jose Antonio 18  

Journal of the International Society of Sports Nutrition volume  7 , Article number:  7 ( 2010 ) Cite this article

533k Accesses

251 Citations

230 Altmetric

Metrics details

Sports nutrition is a constantly evolving field with hundreds of research papers published annually. For this reason, keeping up to date with the literature is often difficult. This paper is a five year update of the sports nutrition review article published as the lead paper to launch the JISSN in 2004 and presents a well-referenced overview of the current state of the science related to how to optimize training and athletic performance through nutrition. More specifically, this paper provides an overview of: 1.) The definitional category of ergogenic aids and dietary supplements; 2.) How dietary supplements are legally regulated; 3.) How to evaluate the scientific merit of nutritional supplements; 4.) General nutritional strategies to optimize performance and enhance recovery; and, 5.) An overview of our current understanding of the ergogenic value of nutrition and dietary supplementation in regards to weight gain, weight loss, and performance enhancement. Our hope is that ISSN members and individuals interested in sports nutrition find this review useful in their daily practice and consultation with their clients.

Introduction

Sports nutrition professionals need to know how to evaluate the scientific merit of articles and advertisements about exercise and nutrition products so they can separate marketing hype from scientifically-based training and nutritional practices. In order to help ISSN members keep informed about the latest in sports nutrition, we have updated the ISSN Exercise & Sports Nutrition Review that was used to help launch the JISSN (originally called the Sports Nutrition Review Journal). This paper provides an overview of: 1.) The definitional category of ergogenic aids and dietary supplements; 2.) How dietary supplements are legally regulated; 3.) How to evaluate the scientific merit of nutritional supplements; 4.) General nutritional strategies to optimize performance and enhance recovery; and, 5.) An overview of our current understanding of the ergogenic value in regards to weight gain, weight loss, and performance enhancement supplements. We have also categorized nutritional supplements into 'apparently effective', 'possibly effective', 'too early to tell', and 'apparently ineffective' as well a description of our general approach into educating athletes about sports nutrition. Over the last five years there have been many changes to our original categorization of supplements. In addition, a number of new supplements have been introduced to the market are reviewed in this article. While some may not agree with all of our interpretations of the literature and/or categorization of a particular supplement, and some classifications may change over time as more research is forthcoming, these interpretations are based on current available scientific evidence and have been well received within the broader scientific community. Our hope is that ISSN members find this information useful in their daily practice and consultation with their clients.

  • Ergogenic Aid

An ergogenic aid is any training technique, mechanical device, nutritional practice, pharmacological method, or psychological technique that can improve exercise performance capacity and/or enhance training adaptations [ 1 – 3 ]. This includes aids that may help prepare an individual to exercise, improve the efficiency of exercise, and/or enhance recovery from exercise. Ergogenic aids may also allow an individual to tolerate heavy training to a greater degree by helping them recover faster or help them stay injury-free and/or healthy during intense training. Although this definition seems rather straightforward, there is considerable debate regarding the ergogenic value of various nutritional supplements. Some sports nutrition specialists only consider a supplement ergogenic if studies show that the supplement significantly enhances exercise performance (e.g., helps you run faster, lift more weight, and/or perform more work during a given exercise task). On the other hand, some feel that if a supplement helps prepare an athlete to perform or enhances recovery from exercise, it has the potential to improve training adaptations and therefore should be considered ergogenic. In the view of the ISSN, one should take a broader view about the ergogenic value of supplements. While we are interested in determining the performance enhancement effects of a supplement on a single bout of exercise, we also realize that one of the goals of training is to help people tolerate a greater degree of training. Individuals who better adapt to high levels of training usually experience greater gains from training over time which can lead to improved performance. Consequently, employing nutritional practices that help prepare individuals to perform and/or enhance recovery from exercise should also be viewed as ergogenic.

Definition and Regulation of Dietary Supplements

As described in Exercise and Sports Nutrition: Principles, Promises, Science & Recommendations [ 3 ]; according to the Food and Drug Administration (FDA), dietary supplements were regulated in the same manner as food prior to 1994 [ 4 ]. Consequently, the FDA monitored the manufacturing processes, quality, and labeling of dietary supplements. However, many people felt that the FDA was too restrictive in regulating dietary supplements. As a result, Congress passed the Dietary Supplement Health and Education Act (DSHEA) in 1994 which placed dietary supplements in a special category of "foods". In October 1994, President Clinton signed DSHEA into law. The law defined a "dietary supplement" as a product taken by mouth that contains a "dietary ingredient" intended to supplement the diet. "Dietary ingredients" may include vitamins, minerals, herbs or other botanicals, amino acids, and substances (e.g., enzymes, organ tissues, glandular, and metabolites). Dietary supplements may also be extracts or concentrates from plants or foods. Dietary supplements are typically sold in the form of tablets, capsules, soft gels, liquids, powders, and bars. Products sold as dietary supplements must be clearly labeled as a dietary supplement.

According to DSHEA, dietary supplements are not drugs. Dietary supplement ingredients that were lawfully sold prior to 1994, have been "grandfathered" into the Act, meaning that a manufacturer is not required to submit to FDA the evidence it relies upon to substantiate safety or effectiveness before or after it markets these ingredients. The rationale for this exclusion is based on a long history of safe use; hence there is no need to require additional safety data. However, DSHEA grants FDA greater control over supplements containing new dietary ingredients. A new dietary ingredient is deemed adulterated and subject to FDA enforcement sanctions unless it meets one of two exemption criteria: either 1.) the supplement in question contains "only dietary ingredients which have been present in the food supply as an article used for food in a form in which the food has not been chemically altered"; or 2.) there is a "history of use or other evidence of safety" provided by the manufacturer or distributor to FDA at least 75 days before introducing the product into interstate commerce. The second criterion, applicable only to new dietary ingredients that have not been present in the food supply, requires manufacturers and distributors of a new dietary ingredient or a product containing a new dietary ingredient to submit pre-market notification to the FDA. This notification, which must be submitted at least 75 days before the product is introduced into interstate commerce, must contain information that provides a history of use or other evidence of safety establishing that the dietary ingredient, when used under the conditions recommended or suggested in the labeling of the dietary supplement will "reasonably be expected to be safe." This may include conducting in vitro toxicology testing, long-term toxicity studies using varying doses in animals to see if there are any toxic effects, providing manufacturing and quality assurance data showing purity, and provision of clinical studies conducted in humans showing safety. The FTC also requires that any representations or claims made about the supplement be substantiated by adequate evidence to show that they are not false or misleading, a policy which is also shared by the FDA. This involves, for example, providing at least two clinical trials showing efficacy of the actual product, within a population of subjects relevant to the target market, supporting the structure/function claims that are made. Structure/function claims may include several categories. They may describe the role of a nutrient or dietary ingredient intended to affect normal structure or function in humans, they may characterize the means by which a nutrient or dietary ingredient acts to maintain such structure or function, they may describe general well-being from consumption of a nutrient or dietary ingredient or they may describe a benefit related to a nutrient deficiency disease, as long as the statement also tells how widespread such a disease is in the United States. Manufacturers of dietary supplements that make structure/function claims on labels or in labeling must submit a notification to FDA no later than 30 days after marketing the dietary supplement that includes the text of the structure/function claim. DSHEA also requires supplement manufacturers to include on any label displaying structure/function claims the disclaimer "This statement has not been evaluated by the FDA. This product is not intended to diagnose, treat, cure, or prevent any disease" . Opponents of dietary supplements often cite this statement as evidence that the FDA did not review or approve the dietary supplement when in fact most dietary ingredients have been grandfathered in due to a long history of safe sale; whereas those products containing a new dietary ingredient which is not present in the food supply as an article used for food in a form in which the food has not been chemically altered are subject to pre-market notification to FDA regarding history of use or other evidence of safety. Unfortunately, a large number of new dietary ingredients requiring pre-market notification have been introduced into dietary supplements since October 1994 without the requisite notification.

According to the 1994 Nutrition Labeling and Education Act (NLEA), the FDA has the ability to review and approve health claims for dietary ingredients and foods. However, since the law was passed it has only approved a few claims. The delay in reviewing health claims of dietary supplements resulted in a lawsuit filed by Pearson & Shaw et al v. Shalala et al in 1993. After years of litigation, the U.S. Court of Appeals for the District of Columbia Circuit ruled in 1999 that qualified health claims may now be made about dietary supplements with approval by FDA as long as the statements are truthful and based on science. Supplement or food companies wishing to make health claims about supplements can submit research evidence to the FDA for approval of a health claim. Additionally, companies must also submit an Investigational New Drug (IND) application to FDA if a research study on a nutrient or multiple dietary ingredient composition is designed to treat an illness and/or medical affliction and/or the company hopes to one day obtain approval for making a qualified health claim as a prescription or orphan drug if the outcome of the study supports the claim. Studies investigating structure/function claims, however, do not need to be submitted to the FDA as an IND. The 1997 Food and Drug Administration Modernization Act (FDAMA) provided for health claims based on an authoritative statement of a scientific body of the U.S. Government or the National Academy of Sciences; such claims may be used after submission of a health claim notification to FDA; and the 2003 FDA Consumer Health Information for Better Nutrition Initiative provided for qualified health claims where the quality and strength of the scientific evidence falls below that required for FDA to issue an authorizing regulation. Such health claims must be qualified to assure accuracy and non-misleading presentation to consumers. More recently, the U.S. Senate passed legislation (Senate Bill 1082) that established the Reagan-Udall Foundation for the FDA. The purpose of this non-profit foundation is to lead collaborations among the FDA, academic research institutions, and industry to enhance research in evaluating the safety and efficacy of dietary supplements as well as to improve the quality and management of these products.

For many years, manufacturers and distributors of dietary supplements were not required to record, investigate or forward to FDA any reports they receive on injuries or illnesses that may be related to the use of their products. However, companies are now required by the Dietary and Supplement and Nonprescription Drug Consumer Act (Public Law 109-462 109th Congress Dec. 22, 2006) to record all adverse event complaints about their products and make them available to the FDA pursuant to an inspection. Reports of "serious" adverse events (i.e., adverse events which results in death, a life-threatening experience, inpatient hospitalization, a persistent or significant disability or incapacity, or a congenital anomaly or birth defect; or requires, based on a reasonable medical judgment, a medical or surgical intervention to prevent an outcome described above) must be reported to FDA within 15 business days. While these reports are unsubstantiated; can be influenced by media attention to a particular supplement; and do not necessarily show a cause and effect: they can be used by the company and FDA to monitor trends and "signals" that may suggest a problem. Once a dietary supplement product is marketed, the FDA has the responsibility for showing that the dietary supplement is unsafe before it can take action to restrict the product's use or removal from the marketplace. The FTC maintains responsibility to make sure manufacturers are truthful and not misleading regarding claims they make about dietary supplements. The FDA has the power to remove supplements from the market if it has sufficient scientific evidence to show the supplement is unsafe. Once they do, they must have sufficient evidence to meet review by the Office of General Accounting (OGA) and/or legal challenges. In the past, the FDA has acted to remove dietary supplements from the market only to be concluded by the OGA and/or federal courts to have overstepped their authority. Additionally, the FTC has the power to act against companies who make false and/or misleading marketing claims about a specific product. This includes acting against companies if the ingredients found in the supplement do not match label claims or in the event undeclared, drug ingredients are present (e.g., analogs of weight loss drugs, diuretic drugs). While this does not ensure the safety of dietary supplements, it does provide a means for governmental oversight of the dietary supplement industry if adequate resources are provided to enforce DSHEA. Since the inception of DSHEA, the FDA has required a number of supplement companies to submit evidence showing safety of their products and acted to remove a number of products sold as dietary supplements from sale in the United States due to safety concerns. Additionally, the FTC has acted against a number of supplement companies for misleading advertisements and/or structure and function claims.

As demonstrated, while some argue that the dietary supplement industry is "unregulated" and/or may have suggestions for additional regulation, manufacturers of dietary supplements must adhere to a number of federal regulations before a product can go to market. Further, they must have evidence that the ingredients sold in their supplements are generally safe if requested to do so by the FDA. For this reason, over the last 20 years, a number of quality supplement companies have employed research and development directors who help educate the public about nutrition and exercise, provide input on product development, conduct preliminary research on products, and/or assist in coordinating research trials conducted by independent research teams (e.g., university based researchers or clinical research sites). They also consult with marketing and legal teams with the responsibility to ensure structure and function claims do not misrepresent results of research findings. This has increased job opportunities for sports nutrition specialists as well as enhanced external funding opportunities for research groups interested in exercise and nutrition research.

While it is true that a number of companies falsely attribute research on different dietary ingredients or dietary supplements to their own, suppress negative findings, and/or exaggerate results from research studies; the trend in the nutrition industry has been to develop scientifically sound supplements. This trend toward greater research support is the result of: 1.) Attempts to honestly and accurately inform the public about results; 2.) Efforts to have data to support safety and efficacy on products for FDA and the FTC; and/or, 3.) To provide scientific evidence to support advertising claims and increase sales. This trend is due in part to greater scrutiny from the FDA and FTC, but also in response to an increasingly competitive marketplace where established safety and efficacy attracts more consumer loyalty and helps ensure a longer lifespan for the product in commerce. In our experience, companies who adhere to these ethical standards prosper while those who do not struggle to comply with FDA and FTC guidelines and rapidly lose consumer confidence, signaling an early demise for the product.

Product Development and Quality Assurance

One of the most common questions raised by athletes, parents, and professionals regarding dietary supplements relates to how they are manufactured and consumer awareness of supplement quality. In a number of cases, reputable companies who develop dietary supplements have research teams who scour the medical and scientific literature looking for potentially effective nutrients. These research teams often attend scientific meetings and review the latest patents, research abstracts presented at scientific meetings, and research publications. They may also consult with leading researchers to discuss ideas about dietary supplements that can be commercialized. Leading companies invest in basic research on nutrients before developing their supplement formulations. Others wait until research has been presented in patents, research abstracts, or publications before developing nutritional formulations featuring the nutrient. Once a new nutrient or formulation has been identified, the next step is to contact raw ingredient suppliers to see if the nutrient can be obtained in a highly pure source and/or if it's affordable. Sometimes, companies develop and patent new processing and purification processes because the nutrient has not yet been extracted in a pure form or is not available in large quantities. Reputable raw material manufacturers conduct extensive tests to examine purity of their raw ingredients. If the company is working on a new ingredient, they often conduct toxicity studies on the new nutrient once a purified source has been identified. They would then compile a safety dossier and communicate it to the FDA as a New Dietary Ingredient submission, with the hopes of it being allowed for lawful sale.

When a powdered formulation is designed, the list of ingredients and raw materials are typically sent to a flavoring house and packaging company to identify the best way to flavor and package the supplement. In the nutrition industry, there are several main flavoring houses and packaging companies who make a large number of dietary supplements for dietary supplement companies. Most reputable dietary supplement manufacturers submit their production facilities to inspection from the FDA and adhere to good manufacturing practices (GMP's), which represent industry standards for good manufacturing of dietary supplements. Some companies also submit their products for independent testing by third-party companies to certify that their products meet label claims. For example, NSF's certification service includes product testing, GMP inspections, ongoing monitoring and use of the NSF Mark indicating products comply with inspection standards, and screening for contaminants. More recently, companies have subjected their products for testing by third party companies to inspect for banned or unwanted substances. These types of tests help ensure that each batch of the dietary supplement does not contained substances banned by the International Olympic Committee or other athletic governing bodies (e.g., NFL). While third-party testing does not guarantee that a supplement is void of banned substances, the likelihood is much less (e.g., Banned Substances Control Group, Informed Choice, etc). Moreover, consumers can request copies of results of these tests. In our experience, companies who are not willing to provide copies of test results are not worth purchasing.

Evaluation of Nutritional Ergogenic Aids

The ISSN recommends going through a process of evaluating the validity and scientific merit of claims made when assessing the ergogenic value of a dietary supplement/technique [ 3 ]. This can be accomplished by examining the theoretical rationale behind the supplement/technique and determining whether there is any well-controlled data showing the supplement/technique works. Supplements based on sound scientific rationale with direct, supportive research showing effectiveness may be worth trying and/or recommending. However, those based on unsound scientific results and/or little to no data supporting the ergogenic value of the actual supplement/technique may not be worthwhile. The sports nutrition specialist should be a resource to help their clients interpret the scientific and medical research that may impact their welfare and/or help them train more wisely and effectively. The following are recommended questions to ask when evaluating the potential ergogenic value of a supplement.

Does The Theory Make Sense?

Most supplements that have been marketed to improve health and/or exercise performance are based on theoretical applications derived from basic and/or clinical research studies. Based on these preliminary studies, a training device or supplement is often marketed to people proclaiming the benefits observed in these basic research studies. Although the theory may appear relevant, critical analysis of this process often reveals flaws in scientific logic and/or that the claims made don't quite match up with the literature cited. By evaluating the literature on your own you can discern whether a supplement has been based on sound scientific evidence or not. To do so, it is suggested you read reviews about the training method, nutrient, and/or supplement from researchers who have been intimately involved in this line of research and/or consult reliable references about nutritional and herbal supplements, such as the JISSN [ 3 , 5 ]. We also suggest doing a search on the nutrient/supplement on the National Library of Medicine's Pub Med Online http://www.ncbi.nlm.nih.gov . A quick look at these references will often help determine if the theory is plausible or not. In our experience, proponents of ergogenic aids often overstate claims made about training devices and/or dietary supplements while opponents of dietary supplements and ergogenic aids are either unaware and/or ignorant of research supporting their use. The sports nutrition specialist has the responsibility to know the literature and/or search available databases to evaluate whether there is merit or not to a proposed ergogenic aid.

Is There Any Scientific Evidence Supporting The Ergogenic Value?

The next question to ask is whether there is any well-controlled data showing effectiveness of the proposed ergogenic aid works as claimed in athletes or people involved in training. The first place to look is the list of references cited in marketing material supporting their claims. We look to see if the abstracts or articles cited are general references or specific studies that have evaluated the efficacy of the nutrient/supplement. We then critically evaluate the abstracts and articles by asking a series of questions.

Are the studies basic research done in animals/clinical populations or have the studies been conducted on athletes/trained subjects? Studies reporting improved performance in rats or persons with type 2 diabetes may be insightful but research conducted on non-diabetic athletes is much more practical and relevant.

Were the studies well controlled? For ergogenic aid research, the study should be a placebo controlled, double-blind, and randomized clinical trial if possible. This means that neither the researcher's nor the subject's were aware which group received the supplement or the placebo during the study and that the subjects were randomly assigned into the placebo or supplement group. An additional element of rigor is called a cross-over design, where each subject, at different times (separated by an interval known as a "washout period"), is exposed to each of the treatments. While utilization of a cross-over design is not always feasible, it removes the element of variability between subjects and increases the strength of the findings. At times, supplement claims have been based on poorly designed studies (i.e., small groups of subjects, no control group, use of unreliable tests, etc) and/or testimonials which make interpretation much more difficult. Well-controlled clinical trials provide stronger evidence as to the potential ergogenic value.

Do the studies report statistically significant results or are claims being made on non-significant means or trends reported? Appropriate statistical analysis of research results allows for an unbiased interpretation of data. Although studies reporting statistical trends may be of interest and lead researchers to conduct additional research, studies reporting statistically significant results are obviously more convincing. With this said, a sports nutrition specialist must be careful not to commit type II statistical errors (i.e., indicating that no differences were observed when a true effect was seen but not detected statistically). Since many studies on ergogenic aids (particularly in high level athletes) evaluate small numbers of subjects, results may not reach statistical significance even though large mean changes were observed. In these cases, additional research is warranted to further examine the potential ergogenic aid before conclusions can be made.

Do the results of the studies cited match the claims made about the supplement? It is not unusual for marketing claims to greatly exaggerate the results found in the actual studies. Additionally, it is not uncommon for ostensibly compelling results, that may indeed by statistically significant, to be amplified while other relevant findings of significant consumer interest are obscured or omitted (e.g. a dietary supplement showing statistically significant increases in circulating testosterone yet changes in body composition or muscular performance were not superior to a placebo). The only way to determine this is to read the entire article, and not just the abstract or even the article citation, and compare results observed in the studies to marketing claims. Reputable companies accurately and completely report results of studies so that consumers can make informed decisions about whether to try a product or not.

Were results of the study presented at a reputable scientific meeting and/or published in a peer-reviewed scientific journal? At times, claims are based on research that has either never been published or only published in an obscure journal. The best research is typically presented at respected scientific meetings and/or published in reputable peer-reviewed journals. Two ways to determine a journal's reputation is either identifying the publisher or the "impact factor" of the journal. A number of "peer-reviewed" journals are published by companies with ties to, or are actually owned by, nutritional products companies (even though they may be available on PubMed). Therefore, we recommend looking up the publisher's website and see how many other journals they publish. If you see only a few other journals this is a suggestion that the journal is not a reputable journal. Alternatively, inquire about the impact factor, a qualitative ranking determined by the number of times a journal's articles are cited. Impact factors are determined and published by Thomson Reuters under Journal Citation Reports ® (a subscription service available at most university libraries). Most journals list their impact factor on the journal home page. The most significant and erudite scientific articles are typically the most read and the most cited.

Have the research findings been replicated at several different labs? The best way to know an ergogenic aid works is to see that results have been replicated in several studies preferably by a number of separate, distinct research groups. The most reliable ergogenic aids are those in which a number of studies, conducted at different labs, have reported similar results of safety and efficacy. Additionally, replication of results by different, unaffiliated labs with completely different authors also removes or reduces the potentially confounding element of publication bias (publication of studies showing only positive results) and conflicts of interest. A notable number of studies on ergogenic aids are conducted in collaboration with one or more research scientists or co-investigators that have a real or perceived economic interest in the outcome of the study. This could range from being a co-inventor on a patent application that is the subject of the ergogenic aid, being paid or receiving royalties from the creation of a dietary supplement formulation, or having stock options or shares in a company that owns or markets the ergogenic aid described in the study. An increasing number of journals require disclosures by all authors of scientific articles, and including such disclosures in published articles. This is driven by the aim of providing greater transparency and research integrity. Disclosure of a conflict of interest does not alone discredit or dilute the merits of a research study. The primary thrust behind public disclosures of potential conflicts of interest is the prevention of a later revelation of an interest that has the potential of discrediting the study in question, the authors, and even the research center or institution where the study was conducted.

Is The Supplement Legal And Safe?

The final question that should be asked is whether the supplement is legal and/or safe. Some athletic associations have banned the use of various nutritional supplements (e.g., prohormones, Ephedra that contains ephedrine, "muscle building" supplements, etc). Obviously, if the supplement is banned, the sports nutrition specialist should discourage its use. In addition, many supplements have not been studied for long-term safety. People who consider taking nutritional supplements should be well aware of the potential side effects so that they can make an informed decision regarding whether to use a supplement or not. Additionally, they should consult with a knowledgeable physician to see if there are any underlying medical problems that may contraindicate use. When evaluating the safety of a supplement, we suggest looking to see if any side effects have been reported in the scientific or medical literature. In particular, we suggest determining how long a particular supplement has been studied, the dosages evaluated, and whether any side effects were observed. We also recommend consulting the Physician's Desk Reference (PDR) for nutritional supplements and herbal supplements to see if any side effects have been reported and/or if there are any known drug interactions. If no side effects have been reported in the scientific/medical literature, we generally will view the supplement as safe for the length of time and dosages evaluated.

Classifying and Categorizing Supplements

Dietary supplements may contain carbohydrate, protein, fat, minerals, vitamins, herbs, enzymes, metabolic intermediates (like amino acids), and/or various plant/food extracts. Supplements can generally be classified as convenience supplements (e.g., energy bars, meal replacement powders, ready to drink supplements) designed to provide a convenient means of meeting caloric needs and/or managing caloric intake, weight gain, weight loss, and/or performance enhancement. Based on the above criteria, we generally categorize nutritional supplements into the following categories:

Apparently Effective . Supplements that help people meet general caloric needs and/or the majority of research studies in relevant populations show is effective and safe.

Possibly Effective . Supplements with initial studies supporting the theoretical rationale but requiring more research to determine how the supplement may affect training and/or performance.

Too Early To Tell . Supplements with sensible theory but lacking sufficient research to support its current use.

Apparently Ineffective . Supplements that lack a sound scientific rationale and/or research has clearly shown to be ineffective.

When a sports nutrition specialist counsels people who train, they should first evaluate their diet and training program. They should make sure that the athlete is eating an energy balanced, nutrient dense diet and that they are training intelligently. This is the foundation to build a good program. Following this, we suggest that they generally only recommend supplements in category I (i.e., 'Apparently Effective). If someone is interested in trying supplements in category II (i.e., 'Possibly Effective'), they should make sure that they understand that these supplements are more experimental and that they may or may not see the type of results claimed. We recommend discouraging people from trying supplements in category III (i.e., 'Too Early to Tell') because there isn't enough data available on their ergogenic value. However, if someone wants to try one of these supplements, they should understand that although there is some theoretical rationale, there is little evidence to support use at this time. Obviously, we do not support athletes taking supplements in categories IV (i.e., 'Apparently Ineffective'). We believe that this approach is a more scientifically supportable and balanced view than simply dismissing the use of all dietary supplements out of hand.

General Dietary Guidelines for Active Individuals

A well-designed diet that meets energy intake needs and incorporates proper timing of nutrients is the foundation upon which a good training program can be developed. Research has clearly shown that not ingesting a sufficient amount of calories and/or enough of the right type of macronutrients may impede an athlete's training adaptations while athletes who consume a balanced diet that meets energy needs can augment physiological training adaptations. Moreover, maintaining an energy deficient diet during training may lead to loss of muscle mass and strength, increased susceptibility to illness, and increased prevalence of overreaching and/or overtraining. Incorporating good dietary practices as part of a training program is one way to help optimize training adaptations and prevent overtraining. The following overviews energy intake and major nutrient needs of active individuals.

Energy Intake

The first component to optimize training and performance through nutrition is to ensure the athlete is consuming enough calories to offset energy expenditure [ 1 , 6 – 8 ]. People who participate in a general fitness program (e.g., exercising 30 - 40 minutes per day, 3 times per week) can typically meet nutritional needs following a normal diet (e.g., 1,800 - 2,400 kcals/day or about 25 - 35 kcals/kg/day for a 50 - 80 kg individual) because their caloric demands from exercise are not too great (e.g., 200 - 400 kcals/session) [ 1 ]. However, athletes involved in moderate levels of intense training (e.g., 2-3 hours per day of intense exercise performed 5-6 times per week) or high volume intense training (e.g., 3-6 hours per day of intense training in 1-2 workouts for 5-6 days per week) may expend 600 - 1,200 kcals or more per hour during exercise [ 1 , 9 ]. For this reason, their caloric needs may approach 50 - 80 kcals/kg/day (2,500 - 8,000 kcals/day for a 50 - 100 kg athlete). For elite athletes, energy expenditure during heavy training or competition may be enormous. For example, energy expenditure for cyclists to compete in the Tour de France has been estimated as high as 12,000 kcals/day (150 - 200 kcals/kg/d for a 60 - 80 kg athlete) [ 9 – 11 ]. Additionally, caloric needs for large athletes (i.e., 100 - 150 kg) may range between 6,000 - 12,000 kcals/day depending on the volume and intensity of different training phases [ 9 ].

Although some argue that athletes can meet caloric needs simply by consuming a well-balanced diet, it is often very difficult for larger athletes and/or athletes engaged in high volume/intense training to be able to eat enough food in order to meet caloric needs [ 1 , 7 , 9 , 10 , 12 ]. Maintaining an energy deficient diet during training often leads to significant weight loss (including muscle mass), illness, onset of physical and psychological symptoms of overtraining, and reductions in performance [ 8 ]. Nutritional analyses of athletes' diets have revealed that many are susceptible to maintaining negative energy intakes during training. Susceptible populations include runners, cyclists, swimmers, triathletes, gymnasts, skaters, dancers, wrestlers, boxers, and athletes attempting to lose weight too quickly [ 7 ]. Additionally, female athletes have been reported to have a high incidence of eating disorders [ 7 ]. Consequently, it is important for the sports nutrition specialist working with athletes to ensure that athletes are well-fed and consume enough calories to offset the increased energy demands of training, and maintain body weight. Although this sounds relatively simple, intense training often suppresses appetite and/or alters hunger patterns so that many athletes do not feel like eating [ 7 ]. Some athletes do not like to exercise within several hours after eating because of sensations of fullness and/or a predisposition to cause gastrointestinal distress. Further, travel and training schedules may limit food availability and/or the types of food athletes are accustomed to eating. This means that care should be taken to plan meal times in concert with training, as well as to make sure athletes have sufficient availability of nutrient dense foods throughout the day for snacking between meals (e.g., drinks, fruit, carbohydrate/protein bars, etc) [ 1 , 6 , 7 ]. For this reason, sports nutritionists' often recommend that athletes consume 4-6 meals per day and snacks in between meals in order to meet energy needs. Use of nutrient dense energy bars and high calorie carbohydrate/protein supplements provides a convenient way for athletes to supplement their diet in order to maintain energy intake during training.

Carbohydrate

The second component to optimizing training and performance through nutrition is to ensure that athletes consume the proper amounts of carbohydrate (CHO), protein (PRO) and fat in their diet. Individuals engaged in a general fitness program can typically meet macronutrient needs by consuming a normal diet (i.e., 45-55% CHO [3-5 grams/kg/day], 10-15% PRO [0.8 - 1.0 gram/kg/day], and 25-35% fat [0.5 - 1.5 grams/kg/day]). However, athletes involved in moderate and high volume training need greater amounts of carbohydrate and protein in their diet to meet macronutrient needs. For example, in terms of carbohydrate needs, athletes involved in moderate amounts of intense training (e.g., 2-3 hours per day of intense exercise performed 5-6 times per week) typically need to consume a diet consisting of 55-65% carbohydrate (i.e., 5-8 grams/kg/day or 250 - 1,200 grams/day for 50 - 150 kg athletes) in order to maintain liver and muscle glycogen stores [ 1 , 6 ]. Research has also shown that athletes involved in high volume intense training (e.g., 3-6 hours per day of intense training in 1-2 workouts for 5-6 days per week) may need to consume 8-10 grams/day of carbohydrate (i.e., 400 - 1,500 grams/day for 50 - 150 kg athletes) in order to maintain muscle glycogen levels [ 1 , 6 ]. This would be equivalent to consuming 0.5 - 2.0 kg of spaghetti. Preferably, the majority of dietary carbohydrate should come from complex carbohydrates with a low to moderate glycemic index (e.g., whole grains, vegetables, fruit, etc). However, since it is physically difficult to consume that much carbohydrate per day when an athlete is involved in intense training, many nutritionists and the sports nutrition specialist recommend that athletes consume concentrated carbohydrate juices/drinks and/or consume high carbohydrate supplements to meet carbohydrate needs.

While consuming this amount of carbohydrate is not necessary for the fitness minded individual who only trains 3-4 times per week for 30-60 minutes, it is essential for competitive athletes engaged in intense moderate to high volume training. The general consensus in the scientific literature is the body can oxidize 1 - 1.1 gram of carbohydrate per minute or about 60 grams per hour [ 13 ]. The American College of Sports Medicine (ACSM) recommends ingesting 0.7 g/kg/hr during exercise in a 6-8% solution (i.e., 6-8 grams per 100 ml of fluid). Harger-Domitrovich et al [ 14 ] reported that 0.6 g/kg/h of maltodextrin optimized carbohydrate utilization [ 14 ]. This would be about 30 - 70 grams of CHO per hour for a 50 - 100 kg individual [ 15 – 17 ]. Studies also indicate that ingestion of additional amounts of carbohydrate does not further increase carbohydrate oxidation.

It should also be noted that exogenous carbohydrate oxidation rates have been shown to differ based on the type of carbohydrate consumed because they are taken up by different transporters [ 18 – 20 ]. For example, oxidation rates of disaccharides and polysaccharides like sucrose, maltose, and maltodextrins are high while fructose, galactose, trehalose, and isomaltulose are lower [ 21 , 22 ]. Ingesting combinations of glucose and sucrose or maltodextrin and fructose have been reported to promote greater exogenous carbohydrate oxidation than other forms of carbohydrate [ 18 – 26 ]. These studies generally indicate a ratio of 1-1.2 for maltodextrin to 0.8-1.0 fructose. For this reason, we recommend that care should be taken to consider the type of carbohydrate to ingest prior to, during, and following intense exercise in order to optimize carbohydrate availability.

There has been considerable debate regarding protein needs of athletes [ 27 – 31 ]. Initially, it was recommended that athletes do not need to ingest more than the RDA for protein (i.e., 0.8 to 1.0 g/kg/d for children, adolescents and adults). However, research over the last decade has indicated that athletes engaged in intense training need to ingest about two times the RDA of protein in their diet (1.5 to 2.0 g/kg/d) in order to maintain protein balance [ 27 , 28 , 30 , 32 , 33 ]. If an insufficient amount of protein is obtained from the diet, an athlete will maintain a negative nitrogen balance, which can increase protein catabolism and slow recovery. Over time, this may lead to muscle wasting and training intolerance [ 1 , 8 ].

For people involved in a general fitness program, protein needs can generally be met by ingesting 0.8 - 1.0 grams/kg/day of protein. Older individuals may also benefit from a higher protein intake (e.g., 1.0 - 1.2 grams/kg/day of protein) in order to help prevent sarcopenia. It is recommended that athletes involved in moderate amounts of intense training consume 1 - 1.5 grams/kg/day of protein (50 - 225 grams/day for a 50 - 150 kg athlete) while athletes involved in high volume intense training consume 1.5 - 2.0 grams/kg/day of protein (75 - 300 grams/day for a 50 - 150 kg athlete) [ 34 ]. This protein need would be equivalent to ingesting 3 - 11 servings of chicken or fish per day for a 50 - 150 kg athlete [ 34 ]. Although smaller athletes typically can ingest this amount of protein in their normal diet, larger athletes often have difficulty consuming this much dietary protein. Additionally, a number of athletic populations have been reported to be susceptible to protein malnutrition (e.g., runners, cyclists, swimmers, triathletes, gymnasts, dancers, skaters, wrestlers, boxers, etc). Therefore, care should be taken to ensure that athletes consume a sufficient amount of quality protein in their diet in order to maintain nitrogen balance (e.g., 1.5 - 2 grams/kg/day).

However, it should be noted that not all protein is the same. Proteins differ based on the source that the protein was obtained, the amino acid profile of the protein, and the methods of processing or isolating the protein [ 35 ]. These differences influence availability of amino acids and peptides that have been reported to possess biological activity (e.g., α-lactalbumin, β-lactoglobulin, glycomacropeptides, immunoglobulins, lactoperoxidases, lactoferrin, etc). Additionally, the rate of digestion and/or absorption and metabolic activity of the protein also are important considerations [ 35 ]. For example, different types of proteins (e.g., casein and whey) are digested at different rates, which directly affect whole body catabolism and anabolism [ 35 – 38 ]. Therefore, care should be taken not only to make sure the athlete consumes enough protein in their diet but also that the protein is high quality. The best dietary sources of low fat, high quality protein are light skinless chicken, fish, egg white and skim milk (casein and whey) [ 35 ]. The best sources of high quality protein found in nutritional supplements are whey, colostrum, casein, milk proteins and egg protein [ 34 , 35 ]. Although some athletes may not need to supplement their diet with protein and some sports nutrition specialists may not think that protein supplements are necessary, it is common for a sports nutrition specialist to recommend that some athletes supplement their diet with protein in order to meet dietary protein needs and/or provide essential amino acids following exercise in order to optimize protein synthesis.

The ISSN has recently adopted a position stand on protein that highlights the following points [ 39 ]:

Exercising individuals need approximately 1.4 to 2.0 grams of protein per kilogram of bodyweight per day.

Concerns that protein intake within this range is unhealthy are unfounded in healthy, exercising individuals.

An attempt should be made to obtain protein requirements from whole foods, but supplemental protein is a safe and convenient method of ingesting high quality dietary protein.

The timing of protein intake in the time period encompassing the exercise session has several benefits including improved recovery and greater gains in fat free mass.

Protein residues such as branched chain amino acids have been shown to be beneficial for the exercising individual, including increasing the rates of protein synthesis, decreasing the rate of protein degradation, and possibly aiding in recovery from exercise.

Exercising individuals need more dietary protein than their sedentary counterparts

The dietary recommendations of fat intake for athletes are similar to or slightly greater than those recommended for non-athletes in order to promote health. Maintenance of energy balance, replenishment of intramuscular triacylglycerol stores and adequate consumption of essential fatty acids are of greater importance among athletes and allow for somewhat increased intake [ 40 ]. This depends on the athlete's training state and goals. For example, higher-fat diets appear to maintain circulating testosterone concentrations better than low-fat diets [ 41 – 43 ]. This has relevance to the documented testosterone suppression which can occur during volume-type overtraining [ 44 ]. Generally, it is recommended that athletes consume a moderate amount of fat (approximately 30% of their daily caloric intake), while increases up to 50% of kcal can be safely ingested by athletes during regular high-volume training [ 40 ]. For athletes attempting to decrease body fat, however, it has been recommended that they consume 0.5 to 1 g/kg/d of fat [ 1 ]. The reason for this is that some weight loss studies indicate that people who are most successful in losing weight and maintaining the weight loss are those who ingest less than 40 g/d of fat in their diet [ 45 , 46 ] although this is not always the case [ 47 ]. Certainly, the type of dietary fat (e.g. n-6 versus n-3; saturation state) is a factor in such research and could play an important role in any discrepancies [ 48 , 49 ]. Strategies to help athletes manage dietary fat intake include teaching them which foods contain various types of fat so that they can make better food choices and how to count fat grams [ 1 , 7 ].

Strategic Eating and Refueling

In addition to the general nutritional guidelines described above, research has also demonstrated that timing and composition of meals consumed may play a role in optimizing performance, training adaptations, and preventing overtraining [ 1 , 6 , 33 , 50 ]. In this regard, it takes about 4 hours for carbohydrate to be digested and begin being stored as muscle and liver glycogen. Consequently, pre-exercise meals should be consumed about 4 to 6 h before exercise [ 6 ]. This means that if an athlete trains in the afternoon, breakfast is the most important meal to top off muscle and liver glycogen levels. Research has also indicated that ingesting a light carbohydrate and protein snack 30 to 60 min prior to exercise (e.g., 50 g of carbohydrate and 5 to 10 g of protein) serves to increase carbohydrate availability toward the end of an intense exercise bout [ 51 , 52 ]. This also serves to increase availability of amino acids and decrease exercise-induced catabolism of protein [ 33 , 51 , 52 ].

When exercise lasts more than one hour, athletes should ingest glucose/electrolyte solution (GES) drinks in order to maintain blood glucose levels, help prevent dehydration, and reduce the immunosuppressive effects of intense exercise [ 6 , 53 – 58 ]. Following intense exercise, athletes should consume carbohydrate and protein (e.g., 1 g/kg of carbohydrate and 0.5 g/kg of protein) within 30 min after exercise as well as consume a high carbohydrate meal within two hours following exercise [ 1 , 31 , 50 ]. This nutritional strategy has been found to accelerate glycogen resynthesis as well as promote a more anabolic hormonal profile that may hasten recovery [ 59 – 61 ]. Finally, for 2 to 3 days prior to competition, athletes should taper training by 30 to 50% and consume 200 to 300 g/d of extra carbohydrate in their diet. This carbohydrate loading technique has been shown to supersaturate carbohydrate stores prior to competition and improve endurance exercise capacity [ 1 , 6 , 50 ]. Thus, the type of meal and timing of eating are important factors in maintaining carbohydrate availability during training and potentially decreasing the incidence of overtraining. The ISSN has a adopted a position stand on nutrient timing [ 13 ] that was summarized with the following points:

Prolonged exercise (> 60 - 90 min) of moderate to high intensity exercise will deplete the internal stores of energy, and prudent timing of nutrient delivery can help offset these changes.

During intense exercise, regular consumption (10 - 15 fl oz.) of a carbohydrate/electrolyte solution delivering 6 - 8% CHO (6 - 8 g CHO/100 ml fluid) should be consumed every 15 - 20 min to sustain blood glucose levels.

Glucose, fructose, sucrose and other high-glycemic CHO sources are easily digested, but fructose consumption should be minimized as it is absorbed at a slower rate and increases the likelihood of gastrointestinal problems.

The addition of PRO (0.15 - 0.25 g PRO/kg/day) to CHO at all time points, especially post-exercise, is well tolerated and may promote greater restoration of muscle glycogen when carbohydrate intakes are suboptimal.

Ingestion of 6 - 20 grams of essential amino acids (EAA) and 30 - 40 grams of high-glycemic CHO within three hours after an exercise bout and immediately before exercise has been shown to significantly stimulate muscle PRO synthesis.

Daily post-exercise ingestion of a CHO + PRO supplement promotes greater increases in strength and improvements in lean tissue and body fat % during regular resistance training.

Milk PRO sources (e.g. whey and casein) exhibit different kinetic digestion patterns and may subsequently differ in their support of training adaptations.

Addition of creatine monohydrate to a CHO + PRO supplement in conjunction with regular resistance training facilitates greater improvements in strength and body composition as compared with when no creatine is consumed.

Dietary focus should center on adequate availability and delivery of CHO and PRO. However, including small amounts of fat does not appear to be harmful, and may help to control glycemic responses during exercise.

Irrespective of timing, regular ingestion of snacks or meals providing both CHO and PRO (3:1 CHO: PRO ratio) helps to promote recovery and replenishment of muscle glycogen when lesser amounts of carbohydrate are consumed.

Vitamins are essential organic compounds that serve to regulate metabolic processes, energy synthesis, neurological processes, and prevent destruction of cells. There are two primary classifications of vitamins: fat and water soluble. The fat soluble vitamins include vitamins A, D, E, & K. The body stores fat soluble vitamins and therefore excessive intake may result in toxicity. Water soluble vitamins are B vitamins and vitamin C. Since these vitamins are water soluble, excessive intake of these vitamins are eliminated in urine, with few exceptions (e.g. vitamin B6, which can cause peripheral nerve damage when consumed in excessive amounts). Table 1 describes RDA, proposed ergogenic benefit, and summary of research findings for fat and water soluble vitamins. Although research has demonstrated that specific vitamins may possess some health benefit (e.g., Vitamin E, niacin, folic acid, vitamin C, etc), few have been reported to directly provide ergogenic value for athletes. However, some vitamins may help athletes tolerate training to a greater degree by reducing oxidative damage (Vitamin E, C) and/or help to maintain a healthy immune system during heavy training (Vitamin C). Theoretically, this may help athletes tolerate heavy training leading to improved performance. The remaining vitamins reviewed appear to have little ergogenic value for athletes who consume a normal, nutrient dense diet. Since dietary analyses of athletes have found deficiencies in caloric and vitamin intake, many sports nutritionists' recommend that athletes consume a low-dose daily multivitamin and/or a vitamin enriched post-workout carbohydrate/protein supplement during periods of heavy training. An article in the Journal of the American Medical Association also recently evaluated the available medical literature and recommended that Americans consume a one-a-day low-dose multivitamin in order to promote general health. Suggestions that there is no benefit of vitamin supplementation for athletes and/or it is unethical for an sports nutrition specialist to recommend that their clients take a one-a-day multi-vitamin and/or suggest taking other vitamins that may raise HDL cholesterol levels and decrease risk of heart disease (niacin), serve as antioxidants (Vitamin E), preserve musculoskeletal function and skeletal mass (vitamin D), or may help maintain a health immune system (Vitamin C) is not consistent with current available literature.

Minerals are essential inorganic elements necessary for a host of metabolic processes. Minerals serve as structure for tissue, important components of enzymes and hormones, and regulators of metabolic and neural control. Some minerals have been found to be deficient in athletes or become deficient in response to training and/or prolonged exercise. When mineral status is inadequate, exercise capacity may be reduced. Dietary supplementation of minerals in deficient athletes has generally been found to improve exercise capacity. Additionally, supplementation of specific minerals in non-deficient athletes has also been reported to affect exercise capacity. Table 2 describes minerals that have been purported to affect exercise capacity in athletes. Of the minerals reviewed, several appear to possess health and/or ergogenic value for athletes under certain conditions. For example, calcium supplementation in athletes susceptible to premature osteoporosis may help maintain bone mass. There is also recent evidence that dietary calcium may help manage body composition. Iron supplementation in athletes prone to iron deficiencies and/or anaemia has been reported to improve exercise capacity. Sodium phosphate loading has been reported to increase maximal oxygen uptake, anaerobic threshold, and improve endurance exercise capacity by 8 to 10%. Increasing dietary availability of salt (sodium chloride) during the initial days of exercise training in the heat has been reported to help maintain fluid balance and prevent dehydration. ACSM recommendations for sodium levels (340 mg) represent the amount of sodium in less than 1/8 teaspoon of salt and meet recommended guidelines for sodium ingestion during exercise (300 - 600 mg per hour or 1.7 - 2.9 grams of salt during a prolonged exercise bout) [ 62 – 65 ]. Finally, zinc supplementation during training has been reported to decrease exercise-induced changes in immune function. Consequently, somewhat in contrast to vitamins, there appear to be several minerals that may enhance exercise capacity and/or training adaptations for athletes under certain conditions. However, although ergogenic value has been purported for remaining minerals, there is little evidence that boron, chromium, magnesium, or vanadium affect exercise capacity or training adaptations in healthy individuals eating a normal diet. Suggestions that there is no benefit of mineral supplementation for athletes and/or it is unethical for a sports nutrition specialist to recommend that their clients take minerals for health and/or performance benefit is not consistent with current available literature.

The most important nutritional ergogenic aid for athletes is water. Exercise performance can be significantly impaired when 2% or more of body weight is lost through sweat. For example, when a 70-kg athlete loses more than 1.4 kg of body weight during exercise (2%), performance capacity is often significantly decreased. Further, weight loss of more than 4% of body weight during exercise may lead to heat illness, heat exhaustion, heat stroke, and possibly death [ 58 ]. For this reason, it is critical that athletes consume a sufficient amount of water and/or GES sports drinks during exercise in order to maintain hydration status. The normal sweat rate of athletes ranges from 0.5 to 2.0 L/h depending on temperature, humidity, exercise intensity, and their sweat response to exercise [ 58 ]. This means that in order to maintain fluid balance and prevent dehydration, athletes need to ingest 0.5 to 2 L/h of fluid in order to offset weight loss. This requires frequent ingestion of 6-8 oz of cold water or a GES sports drink every 5 to 15-min during exercise [ 58 , 66 – 69 ]. Athletes and should not depend on thirst to prompt them to drink because people do not typically get thirsty until they have lost a significant amount of fluid through sweat. Additionally, athletes should weigh themselves prior to and following exercise training to ensure that they maintain proper hydration [ 58 , 66 – 69 ]. The athlete should consume 3 cups of water for every pound lost during exercise in order adequately rehydrate themselves [ 58 ]. Athletes should train themselves to tolerate drinking greater amounts of water during training and make sure that they consume more fluid in hotter/humid environments. Preventing dehydration during exercise is one of the most effective ways to maintain exercise capacity. Finally, inappropriate and excessive weight loss techniques (e.g., cutting weight in saunas, wearing rubber suits, severe dieting, vomiting, using diuretics, etc) are extremely dangerous and should be prohibited. Sports nutrition specialists can play an important role in educating athletes and coaches about proper hydration methods and supervising fluid intake during training and competition.

Dietary Supplements and Athletes

Most of the work we do with athletes regarding sports nutrition is to teach them and their coaches how to structure their diet and time food intake to optimize performance and recovery. Dietary supplements can play a meaningful role in helping athletes consume the proper amount of calories, carbohydrate, and protein in their diet. However, they should be viewed as supplements to the diet, not replacements for a good diet. While it is true that most dietary supplements available for athletes have little scientific data supporting their potential role to enhance training and/or performance, it is also true that a number of nutrients and/or dietary supplements have been shown to help improve performance and/or recovery. Supplementation with these nutrients can help augment the normal diet to help optimize performance. Sports nutrition specialists must be aware of the current data regarding nutrition, exercise, and performance and be honest about educating their clients about results of various studies (whether pro or con). With the proliferation of information available about nutritional supplements to the consumer, the sports nutrition specialist, nutritionist, and nutrition industry lose credibility when they do not accurately describe results of various studies to the public. The following outlines several classifications of nutritional supplements that are often taken by athletes and categorizes them into 'apparently effective', 'possibly effective', 'too early to tell', and 'apparently ineffective' supplements based on interpretation of the literature. It should be noted that this analysis focuses primarily on whether the proposed nutrient has been found to affect exercise and/or training adaptations based on the current available literature. Additional research may or may not reveal ergogenic value, possibly altering its classification. It should be also noted that although there may be little ergogenic value to some nutrients, there may be some potential health benefits that may be helpful for some populations. Therefore, just because a nutrient does not appear to affect performance and/or training adaptations, that does not mean it does not have possible health benefits for athletes.

Convenience Supplements

Convenience supplements are meal replacement powders (MRP's), ready to drink supplements (RTD's), energy bars, and energy gels. They currently represent the largest segment of the dietary supplement industry representing 50 - 75% of most company's sales. They are typically fortified with vitamins and minerals and differ on the amount of carbohydrate, protein, and/or fat they contain. They may also vary based whether they are fortified with various nutrients purported to promote weight gain, enhance weight loss, and/or improve performance. Most people view these supplements as a nutrient dense snack and/or use them to help control caloric intake when trying to gain and/or lose weight. In our view, MRP's, RTD's, and energy bars/gels can provide a convenient way for people to meet specific dietary needs and/or serve as good alternatives to fast food other foods of lower nutritional value. Use of these types of products can be particularly helpful in providing carbohydrate, protein, and other nutrients prior to and/or following exercise in an attempt to optimize nutrient intake when an athlete doesn't have time to sit down for a good meal or wants to minimize food volume. However, they should be used to improve dietary availability of macronutrients - not as a replacement for a good diet. Care should also be taken to make sure they do not contain any banned or prohibited nutrients.

Muscle Building Supplements

The following provides an analysis of the literature regarding purported weight gain supplements and our general interpretation of how they should be categorized based on this information. Table 3 summarizes how we currently classify the ergogenic value of a number of purported performance-enhancing, muscle building, and fat loss supplements based on an analysis of the available scientific evidence.

Apparently Effective

Weight gain powders.

One of the most common means athletes have employed to increase muscle mass is to add extra calories to the diet. Most athletes "bulk up" in this manner by consuming extra food and/or weight gain powders. In order to increase skeletal muscle mass, there must be adequate energy intake (anabolic reactions are endergonic and therefore require adequate energy intake). Studies have consistently shown that simply adding an extra 500 - 1,000 calories per day to your diet in conjunction with resistance training will promote weight gain [ 31 , 33 ]. However, only about 30 - 50% of the weight gained on high calorie diets is muscle while the remaining amount of weight gained is fat. Consequently, increasing muscle mass by ingesting a high calorie diet can help build muscle but the accompanying increase in body fat may not be desirable for everyone. Therefore, we typically do not recommend this type of weight gain approach [ 39 ].

Creatine monohydrate

In our view, the most effective nutritional supplement available to athletes to increase high intensity exercise capacity and muscle mass during training is creatine monohydrate. Numerous studies have indicated that creatine supplementation increases body mass and/or muscle mass during training [ 70 ] Gains are typically 2 - 5 pounds greater than controls during 4 - 12 weeks of training [ 71 ]. The gains in muscle mass appear to be a result of an improved ability to perform high intensity exercise enabling an athlete to train harder and thereby promote greater training adaptations and muscle hypertrophy [ 72 – 75 ]. The only clinically significant side effect occasionally reported from creatine monohydrate supplementation has been the potential for weight gain [ 71 , 76 – 78 ] Although concerns have been raised about the safety and possible side effects of creatine supplementation [ 79 , 80 ], recent long-term safety studies have reported no apparent side effects [ 78 , 81 , 82 ] and/or that creatine monohydrate may lessen the incidence of injury during training [ 83 – 85 ]. Additionally a recent review was published which addresses some of the concerns and myths surrounding creatine monohydrate supplementation [ 86 ]. Consequently, supplementing the diet with creatine monohydrate and/or creatine containing formulations seems to be a safe and effective method to increase muscle mass. The ISSN position stand on creatine monohydrate [ 87 ] summarizes their findings as this:

Creatine monohydrate is the most effective ergogenic nutritional supplement currently available to athletes in terms of increasing high-intensity exercise capacity and lean body mass during training.

Creatine monohydrate supplementation is not only safe, but possibly beneficial in regard to preventing injury and/or management of select medical conditions when taken within recommended guidelines.

There is no compelling scientific evidence that the short- or long-term use of creatine monohydrate has any detrimental effects on otherwise healthy individuals.

If proper precautions and supervision are provided, supplementation in young athletes is acceptable and may provide a nutritional alternative to potentially dangerous anabolic drugs.

At present, creatine monohydrate is the most extensively studied and clinically effective form of creatine for use in nutritional supplements in terms of muscle uptake and ability to increase high-intensity exercise capacity.

The addition of carbohydrate or carbohydrate and protein to a creatine supplement appears to increase muscular retention of creatine, although the effect on performance measures may not be greater than using creatine monohydrate alone.

The quickest method of increasing muscle creatine stores appears to be to consume ~0.3 grams/kg/day of creatine monohydrate for at least 3 days followed by 3-5 g/d thereafter to maintain elevated stores. Ingesting smaller amounts of creatine monohydrate (e.g., 2-3 g/d) will increase muscle creatine stores over a 3-4 week period, however, the performance effects of this method of supplementation are less supported.

Creatine monohydrate has been reported to have a number of potentially beneficial uses in several clinical populations, and further research is warranted in these areas.

As previously described, research has indicated that people undergoing intense training may need additional protein in their diet to meet protein needs (i.e., 1.4 - 2.0 grams/day [ 13 , 39 ] . People who do not ingest enough protein in their diet may exhibit slower recovery and training adaptations [ 33 ]. Protein supplements offer a convenient way to ensure that athletes consume quality protein in the diet and meet their protein needs. However, ingesting additional protein beyond that necessary to meet protein needs does not appear to promote additional gains in strength and muscle mass. The research focus over recent years has been to determine whether different types of protein (e.g., whey, casein, soy, milk proteins, colostrum, etc) and/or various biologically active protein subtypes and peptides (e.g., α-lactalbumin, β-lactoglobulin, glycomacropeptides, immunoglobulins, lactoperoxidases, lactoferrin, etc) have varying effects on the physiological, hormonal, and/or immunological responses to training [ 88 – 91 ]. In addition, a significant amount of research has examined whether timing of protein intake and/or provision of specific amino acids may play a role in protein synthesis and/or training adaptations, conducted mostly in untrained populations [ 92 – 105 ]. Although more research is necessary in this area, evidence clearly indicates that protein needs of individuals engaged in intense training are elevated, different types of protein have varying effects on anabolism and catabolism, that different types of protein subtypes and peptides have unique physiological effects, and timing of protein intake may play an important role in optimizing protein synthesis following exercise. Therefore, it is simplistic and misleading to suggest that there is no data supporting contentions that athletes need more protein in their diet and/or there is no potential ergogenic value of incorporating different types of protein into the diet. It is the position stand of ISSN that exercising individuals need approximately 1.4 to 2.0 grams of protein per kilogram of bodyweight per day. This is greater than the RDA recommendations for sedentary individuals. According to the current literature we know that the addition of protein and or BCAA before or after resistance training can increase protein synthesis and gains in lean mass beyond normal adaptation. However, it should be noted that gains have primarily been observed in untrained populations unless the supplement contained other nutrients like creatine monohydrate [ 13 , 39 ].

Essential Amino Acids (EAA)

Recent studies have indicated that ingesting 3 to 6 g of EAA prior to [ 105 , 106 ] and/or following exercise stimulates protein synthesis [ 92 , 93 , 98 – 101 , 105 ]. Theoretically, this may enhance gains in muscle mass during training. To support this theory, a study by Esmarck and colleagues [ 107 ] found that ingesting EAA with carbohydrate immediately following resistance exercise promoted significantly greater training adaptations in elderly, untrained men, as compared to waiting until 2-hours after exercise to consume the supplement. Although more data is needed, there appears to be strong theoretical rationale and some supportive evidence that EAA supplementation may enhance protein synthesis and training adaptations. Because EAA's include BCAA's, it is probable that positive effects on protein synthesis from EAA ingestion are likely due to the BCAA content [ 108 , 109 ]. Garlick and Grant [ 109 ] infused glucose into growing rats to achieve a concentration of insulin secretion that was insufficient to stimulate protein synthesis by itself. In addition to this, all eight essential amino acids with glucose was infused into another group and then in a third group the investigators only infused the BCAA's along with the glucose. Compared with the glucose infusion alone, protein synthesis was stimulated equally by the essential amino acids and the BCAAs. This demonstrates that the BCAAs are the key amino acids that stimulate protein synthesis. The ISSN position stand on protein concluded that BCAAs have been shown to acutely stimulate protein synthesis, aid in glycogen resynthesis, delaying the onset of fatigue, and help maintain mental function in aerobic-based exercise. It was concluded that consuming BCAAs (in addition to carbohydrates) before, during, and following an exercise bout would be recommended safe and effective [ 39 ].

Possibly Effective

Β-hydroxy β-methylbutyrate (hmb).

HMB is a metabolite of the amino acid leucine. Leucine and metabolites of leucine have been reported to inhibit protein degradation [ 110 ]. Supplementing the diet with 1.5 to 3 g/d of calcium HMB during training has been typically reported to increase muscle mass and strength particularly among untrained subjects initiating training [ 111 – 116 ] and the elderly [ 117 ]. Gains in muscle mass are typically 0.5 to 1 kg greater than controls during 3 - 6 weeks of training. There is also evidence that HMB may lessen the catabolic effects of prolonged exercise [ 118 , 119 ] and that there may be additive effects of co-ingesting HMB with creatine [ 120 , 121 ]. However, the effects of HMB supplementation in athletes are less clear. Most studies conducted on trained subjects have reported non-significant gains in muscle mass possibly due to a greater variability in response of HMB supplementation among athletes [ 122 – 124 ]. Consequently, there is fairly good evidence showing that HMB may enhance training adaptations in individuals initiating training. However, additional research is necessary to determine whether HMB may enhance training adaptations in trained athletes.

Branched Chain Amino Acids (BCAA)

BCAA supplementation has been reported to decrease exercise-induced protein degradation and/or muscle enzyme release (an indicator of muscle damage) possibly by promoting an anti-catabolic hormonal profile [ 31 , 51 , 125 ]. Theoretically, BCAA supplementation during intense training may help minimize protein degradation and thereby lead to greater gains in fat-free mass. There is some evidence to support this hypothesis. For example, Schena and colleagues [ 126 ] reported that BCAA supplementation (~10 g/d) during 21-days of trekking at altitude increased fat free mass (1.5%) while subjects ingesting a placebo had no change in muscle mass. Bigard and associates [ 127 ] reported that BCAA supplementation appeared to minimize loss of muscle mass in subjects training at altitude for 6-weeks. Finally, Candeloro and coworkers [ 128 ] reported that 30 days of BCAA supplementation (14 grams/day) promoted a significant increase in muscle mass (1.3%) and grip strength (+8.1%) in untrained subjects. A recent published abstract [ 129 ] reported that resistance trained subjects ingesting 14 grams of BCAA during 8 weeks of resistance training experienced a significantly greater gain in body weight and lean mass as compared to a whey protein supplemented group and a carbohydrate placebo group. Specifically, the BCAA group gained 2 kg of body mass and 4 kg of lean body mass. In contrast, the whey protein and carbohydrate groups both gained an additional 1 kg of body mass and 2 kg and 1 kg of lean body mass, respectively. It cannot be overstated that this investigation was published as an abstract, was conducted in a gym setting, and has not undergone the rigors of peer review at this time. Although more research is necessary, these findings suggest that BCAA supplementation may have some impact on body composition.

Too Early to Tell

Α-ketoglutarate (α-kg).

α-KG is an intermediate in the Krebs cycle that is involved in aerobic energy metabolism. There is some clinical evidence that α-KG may serve as an anticatabolic nutrient after surgery [ 130 , 131 ]. However, it is unclear whether α-KG supplementation during training may affect training adaptations.

α-Ketoisocaproate (KIC)

KIC is a branched-chain keto acid that is a metabolite of leucine metabolism. In a similar manner as HMB, leucine and metabolites of leucine are believed to possess anticatabolic properties [ 132 ]. There is some clinical evidence that KIC may spare protein degradation in clinical populations [ 133 , 134 ]. Theoretically, KIC may help minimize protein degradation during training possibly leading to greater training adaptations. However, we are not aware of any studies that have evaluated the effects of KIC supplementation during training on body composition.

Ecdysterones

Ecdysterones (also known as ectysterone, 20 Beta-Hydroxyecdysterone, turkesterone, ponasterone, ecdysone, or ecdystene) are naturally derived phytoecdysteroids (i.e., insect hormones). They are typically extracted from the herbs Leuza rhaptonticum sp., Rhaponticum carthamoides , or Cyanotis vaga . They can also be found in high concentrations in the herb Suma (also known as Brazilian Ginseng or Pfaffia). Research from Russia and Czechoslovakia conducted over the last 30 years indicates that ecdysterones may possess some potentially beneficial physiological effects in insects and animals [ 135 – 140 ]. However, since most of the data on ecdysterones have been published in obscure journals, results are difficult to interpret. Wilborn and coworkers [ 141 ] gave resistance trained males 200 mg of 20-hydroxyecdysone per day during 8-weeks of resistance training. It was reported that the 20-hydroxyecdysone supplementation had no effect on fat free mass or anabolic/catabolic hormone status [ 141 ]. Due to the findings of this well controlled study in humans, ecdysterone supplementation at a dosage of 200 mg per day appears to be ineffective in terms of improving lean muscle mass. While future studies may find some ergogenic value of ecdysterones, it is our view that it is too early to tell whether phytoecdysteroids serve as a safe and effective nutritional supplement for athletes.

Growth Hormone Releasing Peptides (GHRP) and Secretagogues

Research has indicated that growth hormone releasing peptides (GHRP) and other non-peptide compounds (secretagogues) appear to help regulate growth hormone (GH) release [ 142 , 143 ]. These observations have served as the basis for development of nutritionally-based GH stimulators (e.g., amino acids, pituitary peptides, "pituitary substances", Macuna pruriens , broad bean, alpha-GPC, etc). Although there is clinical evidence that pharmaceutical grade GHRP's and some non-peptide secretagogues can increase GH and IGF-1 levels at rest and in response to exercise, it has not been demonstrated that such increases lead to an increase in skeletal muscle mass [ 144 ].

Ornithine-α-ketoglutarate (OKG)

OKG (via enteral feeding) has been shown to significantly shorten wound healing time and improve nitrogen balance in severe burn patients [ 145 , 146 ]. Because of its ability to improve nitrogen balance, OKG may provide some value for athletes engaged in intense training. A study by Chetlin and colleagues [ 147 ] reported that OKG supplementation (10 grams/day) during 6-weeks of resistance training promoted greater gains in bench press. However, no significant differences were observed in squat strength, training volume, gains in muscle mass, or fasting insulin and growth hormone. Therefore, additional research is needed before conclusions can be drawn.

Zinc/Magnesium Aspartate (ZMA)

The main ingredients in ZMA formulations are zinc monomethionine aspartate, magnesium aspartate, and vitamin B-6. The rationale of ZMA supplementation is based on studies suggesting that zinc and magnesium deficiency may reduce the production of testosterone and insulin like growth factor (IGF-1). ZMA supplementation has been theorized to increase testosterone and IGF-1 leading to greater recovery, anabolism, and strength during training. In support of this theory, Brilla and Conte [ 148 ] reported that a zinc-magnesium formulation increased testosterone and IGF-1 (two anabolic hormones) leading to greater gains in strength in football players participating in spring training. In another study conducted by Wilborn et al. [ 149 ], resistance trained males ingested a ZMA supplement and found no such increases in either total or free testosterone. In addition, this investigation also assessed changes in fat free mass and no significant differences were observed in relation to fat free mass in those subjects taking ZMA. The discrepancies concerning the two aforementioned studies may be explained by deficiencies of these minerals. Due to the role that zinc deficiency plays relative to androgen metabolism and interaction with steroid receptors [ 150 ], when there are deficiencies of this mineral, testosterone production may suffer. In the study showing increases in testosterone levels [ 148 ], there were depletions of zinc and magnesium in the placebo group over the duration of the study. Hence, increases in testosterone levels could have been attributed to impaired nutritional status rather than a pharmacologic effect. More research is needed to further evaluate the role of ZMA on body composition and strength during training before definitive conclusions can be drawn.

Apparently Ineffective

Glutamine is the most plentiful non-essential amino acid in the body and plays a number of important physiological roles [ 31 , 108 , 109 ] Glutamine has been reported to increase cell volume and stimulate protein [ 151 , 152 ] and glycogen synthesis [ 153 ]. Despite its important role in physiological roles, there is no compelling evidence to support glutamine supplementation in terms of increasing lean body mass. One study that is often cited in support of glutamine supplementation and its role in increasing muscle mass was published by Colker and associates [ 154 ]. It was reported that subjects who supplemented their diet with glutamine (5 grams) and BCAA (3 grams) enriched whey protein during training promoted about a 2 pound greater gain in muscle mass and greater gains in strength than ingesting whey protein alone. While a 2 pound increase in lean body mass was observed, it is likely that these gains were due to the BCAAs that were added to the whey protein. In a well-designed investigation, Candow and co-workers [ 155 ] studied the effects of oral glutamine supplementation combined with resistance training in young adults. Thirty-one participants were randomly allocated to receive either glutamine (0.9 g/kg of lean tissue mass) or a maltodextrin placebo (0.9 g/kg of lean tissue mass) during 6 weeks of total body resistance training. At the end of the 6-week intervention, the authors concluded glutamine supplementation during resistance training had no significant effect on muscle performance, body composition or muscle protein degradation in young healthy adults. While there may be other beneficial uses for glutamine supplementation, there does not appear to be any scientific evidence that it supports increases in lean body mass or muscular performance.

Smilax officinalis (SO)

SO is a plant that contains plant sterols purported to enhance immunity as well as provide an androgenic effect on muscle growth [ 1 ]. Some data supports the potential immune enhancing effects of SO. However, we are not aware of any data that show that SO supplementation increases muscle mass during training.

Isoflavones

Isoflavones are naturally occurring non-steroidal phytoestrogens that have a similar chemical structure as ipriflavone (a synthetic flavonoid drug used in the treatment of osteoporosis) [ 156 – 158 ]. For this reason, soy protein (which is an excellent source of isoflavones) and isoflavone extracts have been investigated in the possible treatment of osteoporosis. Results of these studies have shown promise in preventing declines in bone mass in post-menopausal women as well as reducing risks to side effects associated with estrogen replacement therapy. More recently, the isoflavone extracts 7-isopropoxyisoflavone (ipriflavone) and 5-methyl-7-methoxy-isoflavone (methoxyisoflavone) have been marketed as "powerful anabolic" substances. These claims have been based on research described in patents filed in Hungary in the early 1970s [ 159 , 160 ]. Aubertin-Leheudre M, et al. [ 161 ] investigated the effects that isoflavone supplementation would have on fat-free mass in obese, sarcopenic postmenopausal women. Eighteen sarcopenic-obese women ingested 70 mg of isoflavones per day (44 mg of daidzein, 16 mg glycitein and 10 mg genistein) or a placebo for six months. There was no exercise intervention in the investigation, only the isoflavone supplementation. At the end of the six month intervention, it was reported that there was no difference in total body fat free mass between the isoflavone and placebo groups, but there was a significant increase in the appendicular (arms and legs) fat free mass in the isoflavone supplemented group but not the placebo group. Findings from this study have some applications to sedentary, postmenopausal women. However, there are currently no peer-reviewed data indicating that isoflavone supplementation affects exercise, body composition, or training adaptations in physically active individuals.

Sulfo-Polysaccharides (Myostatin Inhibitors)

Myostatin or growth differentiation factor 8 (GDF-8) is a transforming growth factor that has been shown to serve as a genetic determinant of the upper limit of muscle size and growth [ 162 ]. Recent research has indicated that eliminating and/or inhibiting myostatin gene expression in mice [ 163 ] and cattle [ 164 – 166 ] promotes marked increases in muscle mass during early growth and development. The result is that these animals experience what has been termed as a "double-muscle" phenomenon apparently by allowing muscle to grow beyond its normal genetic limit. In agriculture research, eliminating and/or inhibiting myostatin may serve as an effective way to optimize animal growth leading to larger, leaner, and a more profitable livestock yield. In humans, inhibiting myostatin gene expression has been theorized as a way to prevent or slow down muscle wasting in various diseases, speed up recovery of injured muscles, and/or promote increases in muscle mass and strength in athletes [ 167 ]. While these theoretical possibilities may have great promise, research on the role of myostatin inhibition on muscle growth and repair is in the very early stages - particularly in humans. There is some evidence that myostatin levels are higher in the blood of HIV positive patients who experience muscle wasting and that myostatin levels negatively correlate with muscle mass [ 162 ]. There is also evidence that myostatin gene expression may be fiber specific and that myostatin levels may be influenced by immobilization in animals [ 168 ]. Additionally, a study by Ivey and colleagues [ 167 ] reported that female athletes with a less common myostatin allele (a genetic subtype that may be more resistant to myostatin) experienced greater gains in muscle mass during training and less loss of muscle mass during detraining. No such pattern was observed in men with varying amounts of training histories and muscle mass. These early studies suggest that myostatin may play a role in regulating muscle growth to some degree. Some nutrition supplement companies have marketed sulfo-polysaccharides (derived from a sea algae called Cytoseira canariensis ) as a way to partially bind the myostatin protein in serum. When untrained males supplemented with 1200 mg/day of Cystoseira canariensis in conjunction with a twelve week resistance training regimen, it was reported that there were no differences between the supplemented group and the placebo group in relation to fat-free mass, muscle strength, thigh volume/mass, and serum myostatin [ 169 ]. Interestingly, a recent paper by Seremi and colleagues [ 170 ] reported that resistance training reduced serum myostatin levels and that creatine supplementation in conjunction with resistance training promoted further reductions. Nevertheless, though the research is limited, there is currently no published data supporting the use of sulfo-polysaccharides as a muscle building supplement.

Boron is a trace mineral proposed to increase testosterone levels and promote anabolism. Several studies have evaluated the effects of boron supplementation during training on strength and body composition alterations. These studies (conducted on male bodybuilders) indicate that boron supplementation (2.5 mg/d) appears to have no impact on muscle mass or strength [ 171 , 172 ].

Chromium is a trace mineral that is involved in carbohydrate and fat metabolism. Clinical studies have suggested that chromium may enhance the effects of insulin particularly in diabetic populations. Since insulin is an anti-catabolic hormone and has been reported to affect protein synthesis, chromium supplementation has been theorized to serve as an anabolic nutrient. Theoretically, this may increase anabolic responses to exercise. Although some initial studies reported that chromium supplementation increased gains in muscle mass and strength during training particularly in women [ 173 – 175 ], most well-controlled studies [ 176 ] that have been conducted since then have reported no benefit in healthy individuals taking chromium (200-800 mcg/d) for 4 to 16-weeks during training [ 177 – 183 ]. Consequently, it appears that although chromium supplementation may have some therapeutic benefits for diabetics, chromium does not appear to be a muscle-building nutrient for athletes.

Conjugated Linoleic Acids (CLA)

Animal studies indicate that adding CLA to dietary feed decreases body fat, increases muscle and bone mass, has anti-cancer properties, enhances immunity, and inhibits progression of heart disease [ 184 – 186 ]. Consequently, CLA supplementation in humans has been suggested to help manage body composition, delay loss of bone, and provide health benefit. Although animal studies are impressive [ 187 – 189 ] and some studies suggests benefit over time at some but not all dosages [ 190 – 192 ], there is little current evidence that CLA supplementation during training can affect lean tissue accretion [ 193 , 194 ]. As will be discussed below, there appears to be more promise of CLA as a supplement to promote general health and/or reductions in fat mass over time.

Gamma Oryzanol (Ferulic Acid)

Gamma oryzanol is a plant sterol theorized to increase anabolic hormonal responses during training [ 195 ]. Although data are limited, one study reported no effect of 0.5 g/d of gamma oryzanol supplementation on strength, muscle mass, or anabolic hormonal profiles during 9-weeks of training [ 196 ].

Prohormones

Testosterone and growth hormone are two primary hormones in the body that serve to promote gains in muscle mass (i.e., anabolism) and strength while decreasing muscle breakdown (catabolism) and fat mass [ 197 – 204 ]. Testosterone also promotes male sex characteristics (e.g., hair, deep voice, etc) [ 198 ]. Low level anabolic steroids are often prescribed by physicians to prevent loss of muscle mass for people with various diseases and illnesses [ 205 – 216 ]. It is well known that athletes have experimented with large doses of anabolic steroids in an attempt to enhance training adaptations, increase muscle mass, and/or promote recovery during intense training [ 198 – 200 , 203 , 204 , 217 ]. Research has generally shown that use of anabolic steroids and growth hormone during training can promote gains in strength and muscle mass [ 197 , 202 , 204 , 210 , 213 , 218 – 225 ]. However, a number of potentially life threatening adverse effects of steroid abuse have been reported including liver and hormonal dysfunction, hyperlipidemia (high cholesterol), increased risk to cardiovascular disease, and behavioral changes (i.e., steroid rage) [ 220 , 226 – 230 ]. Some of the adverse effects associated with the use of these agents are irreversible, particularly in women [ 227 ]. For this reason, anabolic steroids have has been banned by most sport organizations and should be avoided unless prescribed by a physician to treat an illness.

Prohormones (androstenedione, 4-androstenediol, 19-nor-4-androstenedione, 19-nor-4-androstenediol, 7-keto DHEA, and DHEA, etc) are naturally derived precursors to testosterone or other anabolic steroids. Prohormones have become popular among body builders because they believe they are natural boosters of anabolic hormones. Consequently, a number of over-the-counter supplements contain prohormones. While there is some data indicating that prohormones increase testosterone levels [ 231 , 232 ], there is virtually no evidence that these compounds affect training adaptations in younger men with normal hormone levels. In fact, most studies indicate that they do not affect testosterone and that some may actually increase estrogen levels and reduce HDL-cholesterol [ 220 , 231 , 233 – 238 ]. Consequently, although there may be some potential applications for older individuals to replace diminishing androgen levels, it appears that prohormones have no training value. Since prohormones are "steroid-like compounds", most athletic organizations have banned their use. Use of nutritional supplements containing prohormones will result in a positive drug test for anabolic steroids. Use of supplements knowingly or unknowingly containing prohormones have been believed to have contributed to a number of recent positive drug tests among athletes. Consequently, care should be taken to make sure that any supplement an athlete considers taking does not contain prohormone precursors particularly if their sport bans and tests for use of such compounds. It is noteworthy to mention that many prohormones are not lawful for sale in the USA since the passage of the Anabolic Steroid Control Act of 2004. The distinctive exception to this is DHEA, which has been the subject of numerous clinical studies in aging populations.

Rather than provide the body with a precursor to testosterone, a more recent technique to enhance endogenous testosterone has been to inhibit aromatase activity [ 239 ]. Two studies have investigated the effects of aromatase inhibitors (androst-4-ene-3,6,17-trione) [ 240 ] and (hydroxyandrost-4-ene-6,17-dioxo-3-THP ether and 3,17-diketo-androst-1,4,6-triene) [ 241 ]. In both of these investigations, it was reported that free testosterone and dihydrotesterone levels were significantly increased. Muscle mass/fat free mass was not measured in one investigation [ 240 ] and no changes were observed in fat free mass in the other investigation [ 241 ].

Tribulus terrestris

Tribulus terrestris (also known as puncture weed/vine or caltrops) is a plant extract that has been suggested to stimulate leutinizing hormone (LH) which stimulates the natural production of testosterone [ 132 ]. Consequently, Tribulus has been marketed as a supplement that can increase testosterone and promote greater gains in strength and muscle mass during training. Several recent studies have indicated that Tribulus supplementation appears to have no effects on body composition or strength during training [ 242 – 244 ].

Vanadyl Sulfate (Vanadium)

In a similar manner as chromium, vanadyl sulfate is a trace mineral that has been found to affect insulin-sensitivity and may affect protein and glucose metabolism [ 132 , 245 ]. For this reason, vanadyl sulfate has been purported to increase muscle mass and strength during training. Although there may be some clinical benefits for diabetics (with a therapeutic dose of at least 50 mg vanadyl sulfate twice daily [ 246 , 247 ], vanadyl sulfate supplementation does not appear to have any effect on strength or muscle mass during training in non-diabetic, weight training individuals [ 248 , 249 ].

Weight Loss Supplements

Although exercise and proper diet remain the best way to promote weight loss and/or manage body composition, a number of nutritional approaches have been investigated as possible weight loss methods (with or without exercise). The following overviews the major types of weight loss products available and discusses whether any available research supports their use. See Table 3 for a summary.

Low Calorie Diet Foods & Supplements

Most of the products in this category represent low fat/carbohydrate, high protein food alternatives [ 250 ]. They typically consist of pre-packaged food, bars, MRP, or RTD supplements. They are designed to provide convenient foods/snacks to help people follow a particular low calorie diet plan. In the scientific literature, diets that provide less than 1000 calories per day are known as very low calorie diets (VLCD's). Pre-packaged food, MRP's, and/or RTD's are often provided in VLCD plans to help people cut calories. In most cases, VLCD plans recommend behavioural modification and that people start a general exercise program.

Research on the safety and efficacy of people maintaining VLCD's generally indicate that they can promote weight loss. For example, Hoie et al [ 251 ] reported that maintaining a VLCD for 8-weeks promoted a 27 lbs (12.6%) loss in total body mass, a 21 lbs loss in body fat (23.8%), and a 7 lbs (5.2%) loss in lean body mass in 127 overweight volunteers. Bryner and colleagues [ 252 ] reported that addition of a resistance training program while maintaining a VLCD (800 kcal/d for 12-weeks) resulted in a better preservation of lean body mass and resting metabolic rate compared to subjects maintaining a VLCD while engaged in an endurance training program. Meckling and Sherfey [ 253 ] reported that the combination of high protein and exercise was the most effective intervention for weight loss and was superior to a low-fat, high-carbohydrate diet in promoting weight loss and nitrogen balance regardless of the presence of an exercise intervention. Recent studies indicate that high protein/low fat VLCD's may be better than high carbohydrate/low fat diets in promoting weight loss [ 46 , 253 – 260 ]. The reason for this is that typically when people lose weight about 40-50% of the weight loss is muscle which decreases resting energy expenditure. Increasing protein intake during weight loss helps preserve muscle mass and resting energy expenditure to a better degree than high carbohydrate diets [ 261 , 262 ]. These findings and others indicate that VLCD's (typically using MRP's and/or RTD's as a means to control caloric intake) can be effective particularly as part of an exercise and behavioural modification program. Most people appear to maintain at least half of the initial weight lost for 1-2 years but tend to regain most of the weight back within 2-5 years. Therefore, although these diets may help people lose weight on the short-term, it is essential people who use them follow good diet and exercise practices in order to maintain the weight loss. The addition of dietary protein whether in whole food form or meal replacement form could assist in this weight maintenance due to the fact that the retention of muscle mass is greater than in high carbohydrate/low-fat weight loss trials?

Ephedra, Caffeine, and Silicin

Thermogenics are supplements designed to stimulate metabolism thereby increasing energy expenditure and promote weight loss. They typically contain the "ECA" stack of ephedra alkaloids (e.g., Ma Haung, 1R,2S Nor-ephedrine HCl, Sida Cordifolia), caffeine (e.g., Gaurana, Bissey Nut, Kola) and aspirin/salicin (e.g., Willow Bark Extract). The first of the three traditional thermogenics is now banned by the FDA however the safety associated with the ingestion of ephedra is debated. More recently, other potentially thermogenic nutrients have been added to various thermogenic formulations. For example, thermogenic supplements may also contain synephrine (e.g., Citrus Aurantum, Bitter Orange), calcium & sodium phosphate, thyroid stimulators (e.g., guggulsterones, L-tyrosine, iodine), cayenne & black pepper, and ginger root.

A significant amount of research has evaluated the safety and efficacy of EC and ECA type supplements. According to a meta-analysis in the Journal of American Medical Association, ephedrine/ephedra promote a more substantial weight loss 0.9 kg per month in comparison to placebo in clinical trials but are associated with increased risk of psychiatric, autonomic or gastrointestinal symptoms as well as heart palpitations. Several studies have confirmed that use of synthetic or herbal sources of ephedrine and caffeine (EC) promote about 2 lbs of extra weight loss per month while dieting (with or without exercise) and that EC supplementation is generally well tolerated in healthy individuals [ 263 – 274 ]. For example, Boozer et al [ 267 ] reported that 8-weeks of ephedrine (72 mg/d) and caffeine (240 mg/d) supplementation promoted a 9 lbs loss in body mass and a 2.1% loss in body fat with minor side effects. Hackman and associates [ 275 ] reported that a 9 month clinical trial utilizing a multi-nutrient supplement containing 40 mg/d of ephedra alkaloids and 100 mg/day caffeine resulted in a loss of weight and body fat, improved metabolic parameters including insulin sensitivity without any apparent side effects. Interestingly, Greenway and colleagues [ 274 ] reported that EC supplementation was a more cost-effective treatment for reducing weight, cardiac risk, and LDL cholesterol than several weight loss drugs (fenfluramine with mazindol or phentermine). Finally, Boozer and associates [ 268 ] reported that 6-months of herbal EC supplementation promoted weight loss with no clinically significant adverse effects in healthy overweight adults. Less is known about the safety and efficacy of synephrine, thyroid stimulators, cayenne/black pepper and ginger root.

Despite these findings, the Food and Drug Administration (FDA) banned the sale of ephedra containing supplements. The rationale has been based on reports to adverse event monitoring systems and in the media suggesting a link between intake of ephedra and a number of severe medical complications (e.g., high blood pressure, elevated heart rate, arrhythmias, sudden death, heat stroke, etc) [ 276 , 277 ]. Although results of available clinical studies do not show these types of adverse events, ephedra is no longer available as an ingredient in dietary supplements and thus cannot be recommended for use. Consequently, thermogenic supplements now contain other nutrients believed to increase energy expenditure (e.g., synephrine, green tea, etc) and are sold as "ephedrine-free" types of products. Anyone contemplating taking thermogenic supplements should carefully consider the potential side effects, discuss possible use with a knowledgeable physician, and be careful not to exceed recommended dosages.

High Fiber Diets

One of the oldest and most common methods of suppressing the appetite is to consume a diet that is high in fiber. Ingesting high fiber foods (fruits, vegetables) or fiber containing supplements (e.g., glucomannan) increase the feeling of fullness (satiety) which typically allows an individual to feel full while ingesting fewer calories. Theoretically, maintaining a high fiber diet may serve to help decrease the amount of food you eat. In addition, high fiber diets/supplements help lower cholesterol and blood pressure, enhance insulin sensitivity, and promote weight loss in obese subjects [ 278 ]. A recent study found that a Mediterranean diet that was high in fiber resulted in a more dramatic weight loss that a traditional low-fat diet and had beneficial effects on glycemic control [ 279 ]. Other research on high fiber diets indicates that they provide some benefit, particularly in diabetic populations. For example, Raben et al [ 280 ] reported that subjects maintaining a low fat/high fiber diet for 11 weeks lost about 3 lbs of weight and 3.5 lbs of fat. Other studies have reported mixed results on altering body composition using various forms of higher fiber diets [ 281 – 284 ]. Consequently, although maintaining a low fat/high fiber diet that is high in fruit and vegetable content has various health benefits, these diets seem to have potential to promote weight loss as well as weight maintenance thus we can recommend high fiber diets as a safe and healthy approach to possibly improve body composition.

Several studies and recent reviews have reported that calcium supplementation alone or in combination with other ingredients does not affect weight loss or fat loss [ 285 – 290 ]. Research has indicated that calcium modulates 1,25-diydroxyvitamin D which serves to regulate intracellular calcium levels in fat cells [ 291 , 292 ]. Increasing dietary availability of calcium reduces 1,25-diydroxyvitamin D and promotes reductions in fat mass in animals [ 292 – 294 ]. Dietary calcium has been shown to suppress fat metabolism and weight gain during periods of high caloric intake [ 291 , 293 , 295 ]. Further, increasing calcium intake has been shown to increase fat metabolism and preserve thermogenesis during caloric restriction [ 291 , 293 , 295 ]. In support of this theory, Davies and colleagues [ 296 ] reported that dietary calcium was negatively correlated to weight and that calcium supplementation (1,000 mg/d) accounted for an 8 kg weight loss over a 4 yr period. Additionally, Zemel and associates [ 291 ] reported that supplemental calcium (800 mg/d) or high dietary intake of calcium (1,200 - 1,300 mg/d) during a 24-week weight loss program promoted significantly greater weight loss (26-70%) and dual energy x-ray absorptiometer (DEXA) determined fat mass loss (38-64%) compared to subjects on a low calcium diet (400-500 mg/d). These findings and others suggest a strong relationship between calcium intake and fat loss. However, more research needs to be conducted before definitive conclusions can be drawn.

Green Tea Extract

Green tea is now one of the most common herbal supplements that is being added to thermogenic products because it has been suggested to affect weight loss and is now the fourth most commonly used dietary supplement in the US [ 297 ]. Green tea contains high amounts of caffeine and catechin polyphenols. The primary catechin that is associated to the potential effects on weight loss through diet induced thermogenesis is the catechin epigallocatechin gallate, also known as EGCG [ 298 , 299 ]. Research suggests that catechin polyphenols possess antioxidant properties and the intake of tea catechins is associated with a reduced risk of cardiovascular disease [ 298 – 300 ]. In addition, green tea has also been theorized to increase energy expenditure by stimulating brown adipose tissue thermogenesis. In support of this theory, Dulloo et al [ 301 , 302 ] reported that green tea supplementation in combination with caffeine (e.g., 50 mg caffeine and 90 mg epigallocatechin gallate taken 3-times per day) significantly increased 24-hour energy expenditure and fat utilization in humans to a much greater extent than when an equivalent amount of caffeine was evaluated suggesting a synergistic effect. Recently, work by Di Pierro and colleagues [ 303 ] reported that the addition of a green tea extract to a hypocaloric diet resulted in a significant increase in weight loss (14 kg vs. 5 kg) versus a hypocaloric diet alone over a 90 day clinical trial. Maki and coworkers [ 304 ] also demonstrated that green tea catechin consumption enhanced the exercise-induced changes in abdominal fat. However, it must be noted that both human and animal studies have not supported these findings and have reported that supplementation of these extracts does not affect weight loss [ 305 , 306 ]. Theoretically, increases in energy expenditure may help individuals lose weight and/or manage body composition.

CLA is a term used to describe a group of positional and geometric isomers of linoleic acid that contain conjugated double bonds. Adding CLA to the diet has been reported to possess significant health benefits in animals [ 184 , 307 ]. In terms of weight loss, CLA feedings to animals have been reported to markedly decrease body fat accumulation [ 185 , 308 ]. Consequently, CLA has been marketed as a health and weight loss supplement since the mid 1990s. Despite the evidence in animal models, the effect of CLA supplementation in humans is less clear. There are some data suggesting that CLA supplementation may modestly promote fat loss and/or increases in lean mass [ 190 – 192 , 309 – 314 ]. Recent work suggested that CLA supplementation coupled with creatine and whey protein resulted in a increase in strength and lean-tissue mass during resistance training [ 315 ]. However, other studies indicate that CLA supplementation (1.7 to 12 g/d for 4-weeks to 6-months) has limited to no effects on body composition alterations in untrained or trained populations [ 190 , 310 , 316 – 324 ]. The reason for the discrepancy in research findings has been suggested to be due to differences in purity and the specific isomer studied. For instance, early studies in humans showing no effect used CLA that contained all 24 isomers. Today, most labs studying CLA use 50-50 mixtures containing the trans-10, cis-12 and cis-9, trans-11 isomers, the former of which being recently implicated in positively altering body composition. This has been supported by recent work indicating that CLA (50:50 cis-9, trans-11:trans-10, cis-12) plus polyunsaturated fatty acid supplementation prevented abdominal fat increases and increase fat-free mass [ 325 ]. However, it must be noted that this response only occurred in young obese individuals. Thus, CLA supplementation may have potential in the areas of general health and it is clear that research on the effects on body composition is ongoing and still quite varied. Further research is needed to determine which CLA isomer is ideal for ingestion and possibly if there are differential responses among lean or obese and old or young populations.

Gymnema Sylvestre

Gymnema Sylvestre is a supplement that is purported to regulate weight loss and blood sugar levels. It is purported to affect glucose and fat metabolism as well as inhibit sweet cravings. In support of these contentions, some recent data have been published by Shigematsu and colleagues [ 326 , 327 ] showing that short and long-term oral supplementation of gymnema sylvestre in rats fed normal and high-fat diets may have some positive effects on fat metabolism, blood lipid levels, and/or weight gain/fat deposition. More recent work in rats has shown that gymnema sylvestre supplementation promoted weight loss by reducing hyperlipidemia [ 328 ]. The only apparent clinical trial in humans showed that an herbal combination group containing 400 mg of gymnema sylvestra resulted in effective and safe weight loss while promoting improved blood lipid profiles. It should be noted that this group was not significantly different that the other active group, containing HCA, when observing these dependent variables [ 329 ]. Due to the lack of substantial positive research on the effects of gymnema sylvestre supplementation in humans, we cannot recommend gymnema sylvestre as a supplement to positively affect weight loss.

Phosphatidyl Choline (Lecithin)

Choline is considered an essential nutrient that is needed for cell membrane integrity and to facilitate the movement of fats in and out of cells. It is also a component of the neurotransmitter acetylcholine and is needed for normal brain functioning, particularly in infants. For this reason, phosphatidyl choline (PC) has been purported as a potentially effective supplement to promote fat loss as well as improve neuromuscular function. However, despite these alleged benefits of lecithin supplementation, there are no clinical trials in humans to support a potential role of lecithin supplementation affecting weight loss.

Betaine is a compound that is involved in the metabolism of choline and homocysteine. Garcia Neto et al. [ 330 ] have shown that betaine feedings can effect liver metabolism, fat metabolism, and fat deposition in chickens. Betaine supplementation may also help lower homocysteine levels which is a marker of risk to heart disease [ 331 ]. For this reason, betaine supplements have been marketed as a supplement designed to promote heart health as well as a weight loss. A recent study by Hoffman and colleagues [ 332 ] found betaine supplementation to improve muscular endurance in active college age males. Despite this, there appears to be little evidence in human models that supports the role of betaine as a supplement for weight loss and thus it is not recommended for supplementation.

Coleus Forskohlii (Forskolin)

Forskolin, which is touted as a weight loss supplement is a plant native to India that has been used for centuries in traditional Ayurvedic medicine primarily to treat skin disorders and respiratory problems [ 333 , 334 ]. A considerable amount of research has evaluated the physiological and potential medical applications of forskolin over the last 25 years. Forskolin has been reported to reduce blood pressure, increase the hearts ability to contract, help inhibit platelet aggregation, improve lung function, and aid in the treatment of glaucoma [ 333 – 335 ]. With regard to weight loss, forskolin has been reported to increase cyclic AMP and thereby stimulate fat metabolism [ 336 – 338 ]. Theoretically, forskolin may therefore serve as an effective weight loss supplement. Recent evidence has shown that forskolin supplementation had no effect on improving body composition in mildly obese women [ 339 ]. In contrast, work done by Godard et al. in 2005 reported that 250 mg of a 10% forskolin extract taken twice daily resulted in improvements in body composition in overweight and obese men [ 340 ]. Another study suggested that supplementing the diet with coleus forskohlii in overweight women helped maintain weight and was not associated with any clinically significant adverse events [ 341 ]. Currently, research is still needed on forskolin supplementation before it can be recommended as an effective weight loss supplement.

Dehydroepiandrosterone (DHEA) and 7-Keto DHEA

Dehydroepiandrosterone (DHEA) and its sulfated conjugate DHEAS represent the most abundant adrenal steroids in circulation [ 342 ]. Although, DHEA is considered a weak androgen, it can be converted to the more potent androgens testosterone and dihydrotestosterone in tissues. In addition, DHEAS can be converted into androstenedione and testosterone. DHEA levels have been reported to decline with age in humans [ 343 ]. The decline in DHEA levels with aging has been associated with increased fat accumulation and risk to heart disease [ 344 ]. Since DHEA is a naturally occurring compound, it has been suggested that dietary supplementation of DHEA may help maintain DHEA availability, maintain and/or increase testosterone levels, reduce body fat accumulation, and/or reduce risk to heart disease as one ages [ 342 , 344 ]. Although animal studies have generally supported this theory, the effects of DHEA supplementation on body composition in human trials have been mixed. For example, Nestler and coworkers [ 345 ] reported that DHEA supplementation (1,600 mg/d for 28-d) in untrained healthy males promoted a 31% reduction in percentage of body fat. However, Vogiatzi and associates [ 346 ] reported that DHEA supplementation (40 mg/d for 8 wks) had no effect on body weight, percent body fat, or serum lipid levels in obese adolescents. More recent work has supported these findings suggesting that one year of DHEA supplementation had no effect on body composition when taken at 50 mg per day [ 347 ]. 7-keto DHEA, a DHEA precursor, has been marketed as a potentially more effective form of DHEA which is believed to possess lypolytic properties. Although data are limited, Kalman and colleagues and coworkers [ 348 ] reported that 7-keto DHEA supplementation (200 mg/d) during 8-weeks of training promoted a greater loss in body mass and fat mass while increasing T3 while observing no significant effects on thyroid stimulating hormone (TSH) or T4. More recent data has shown that 7-keto DHEA supplementation can increase RMR [ 349 ] and blunt the decrease in RMR associated with 8 weeks of restricted dieting [ 350 ]. However, it must be noted that the second study did not use isolated 7-keto DHEA but used a commercial weight loss product that contained DHEA as well as other known weight loss agents (i.e. caffeine, green tea extract, citrus aurantium, etc.). Thus, these results do not directly support the use of 7-keto DHEA. Although more research is needed on the effects of supplementing DHEA by itself as a weight loss agent, these findings provide minimal support that 7-keto DHEA may serve as an effective weight loss supplement.

Psychotropic Nutrients/Herbs

Psychotropic nutrients/herbs are a new class of supplements that often contain things like St. John's Wart, Kava, Ginkgo Biloba, Ginseng, and L-Tyrosine. They are believed to serve as naturally occurring antidepressants, relaxants, and mental stimulants thus the theoretical rationale regarding weight loss is that they may help people fight depression or maintain mental alertness while dieting. There are no clinical weight loss trials that utilize any of the above nutrients/herbs as the active ingredient in the supplementation trial. Although a number of studies support potential role as naturally occurring psychotropics or stimulants, the potential value in promoting weight loss is unclear and therefore are not recommended for supplementation.

Calcium Pyruvate

Calcium Pyruvate is supplement that hit the scene about 10-15 ago with great promise. The theoretical rationale was based on studies from the early 1990s that reported that calcium pyruvate supplementation (16 - 25 g/d with or without dihydroxyacetone phosphate [DHAP]) promoted fat loss in overweight/obese patients following a medically supervised weight loss program [ 351 – 353 ]. Although the mechanism for these findings was unclear, the researchers speculated that it might be related to appetite suppression and/or altered carbohydrate and fat metabolism. Since calcium pyruvate is very expensive, several studies have attempted to determine whether ingesting smaller amounts of calcium pyruvate (6-10 g/d) affect body composition in untrained and trained populations. Results of these studies are mixed. Earlier studies have shown both a positive effect on calcium pyruvate supplementation in improving body composition [ 354 ], however, Stone and colleagues [ 355 ] reported that pyruvate supplementation did not affect hydrostatically determined body composition during 5-weeks of in-season college football training. More recently, calcium pyruvate supplementation was also shown to not have a significant effect on body composition or exercise performance. Additionally, it has been reported that supplementation may negatively affect some blood lipid levels [ 356 ]. These findings indicate that although there is some supportive data indicating that calcium pyruvate supplementation may enhance fat loss when taken at high doses (6-16 g/d), there is no evidence that ingesting the doses typically found in pyruvate supplements (0.5 - 2 g/d) has any affect on body composition. In addition, the overall quantity of research examining calcium pyruvate is minimal at best thus it is not warranted to include calcium pyruvate as a weight loss supplement.

Chitosan has been marketed as a weight loss supplement for several years as is known as a "fat trapper". It is purported to inhibit fat absorption and lower cholesterol. This notion is supported animal studies indicated by decreased fat absorption, increased fat content, and/or lower cholesterol following chitosan feedings [ 357 – 360 ]. However, the effects in humans appear to be less impressive. For example, although there is some data suggesting that chitosan supplementation may lower blood lipids in humans,[ 361 ] other studies report no effects on fat content [ 362 , 363 ]or body composition alterations [ 364 – 366 ] when administered to people following their normal diet. More recent work has shown that the effect of chitosan on fat absorption is negligible and is the equivalent of approximately 9.9 kcal/day following supplementation [ 362 ]. Other work has concluded that the insignificant amounts of fat that are trapped from supplementation would take about 7 months for a male to lose a pound of weight, and that the effect was completely ineffective in women [ 364 ]. Thus, based on the current evidence, chitosan supplementation is apparently ineffective and has no significant effects on "fat trapping" and/or on improving body composition.

Chromium supplementation is derived from its role in maintaining proper carbohydrate and fat metabolism by potentially effecting insulin signalling [ 367 ]. Initial studies reported that chromium supplementation during resistance training improved fat loss and gains in lean body mass [ 173 – 175 ]. To date, the studies using more accurate methods of assessing body composition have primarily indicate no effects on body composition in healthy non-diabetic individuals [ 176 – 183 , 368 ]. Recent work has reported that 200 mcg of chromium picolinate supplementation on individuals on a restrictive diet did not promote weight loss or body composition changes following 12 weeks of supplementation [ 368 ]. This work supports Lukaski et al [ 182 ] previous findings that 8-weeks of chromium supplementation during resistance training did not affect strength or DEXA determined body composition changes. Thus, based on the current review of the literature we cannot recommend chromium supplementation as a means of improving body composition.

Garcinia Cambogia (HCA)

HCA is a nutrient that has been hypothesized to increase fat oxidation by inhibiting citrate lypase and lipogenesis [ 369 ]. Theoretically, this may lead to greater fat burning and weight loss over time. Although there is some evidence that HCA may increase fat metabolism in animal studies, there is little to no evidence showing that HCA supplementation affects body composition in humans. For example, Ishihara et al [ 370 ] reported that HCA supplementation spared carbohydrate utilization and promoted lipid oxidation during exercise in mice. However, Kriketos and associates [ 371 ] reported that HCA supplementation (3 g/d for 3-days) did not affect resting or post-exercise energy expenditure or markers of lipolysis in healthy men. Likewise, Heymsfield and coworkers [ 372 ] reported that HCA supplementation (1.5 g/d for 12-weeks) while maintaining a low fat/high fiber diet did not promote greater weight or fat loss than subjects on placebo. Finally, Mattes and colleagues [ 373 ] reported that HCA supplementation (2.4 g/d for 12-weeks) did not affect appetite, energy intake, or weight loss. These findings suggest that HCA supplementation does not appear to promote fat loss in humans.

L-Carnitine

Carnitine serves as an important transporter of fatty acids from the cytosol into the mitochondria of the cell [ 374 ]. Increased cellular levels of carnitine would theoretically enhance transport of fats into the mitochondria and thus provide more substrates for fat metabolism. L-carnitine has been one of the most common nutrients found in various weight loss supplements. Over the years, a number of studies have been conducted on the effects of L-carnitine supplementation on fat metabolism, exercise capacity and body composition. The overwhelming conclusions of L-carnitine research indicates that L-carnitine supplementation does not affect muscle carnitine content [ 375 ], fat metabolism, aerobic- or anaerobic-exercise performance [ 375 ], and/or weight loss in overweight or trained subjects [ 376 , 377 ]. Despite the fact that L-carnitine has been shown apparently ineffective as a supplement, the research on L-carnitine has shifted to another category revolving around hypoxic stress and oxidative stress. Preliminary research has reported that L-carnitine supplementation has a minimal effect on reducing the biomarkers of exercise-induced oxidative stress [ 378 ]. While these findings are not promising, there is some recent data indicating that L-carnitine tartrate supplementation during intensified periods of training may help athletes tolerate training to a greater degree [ 379 ]. Consequently, there may be other advantages to L-carnitine supplementation than promoting fat metabolism.

The role of sodium and calcium phosphate on energy metabolism and exercise performance has been studied for decades [ 31 ]. Phosphate supplementation has also been suggested to affect energy expenditure, however, the research in this area is quite dated and no research on the effects on energy expenditure have been conducted. Some of this dated work includes the work by Kaciuba-Uscilko and colleagues [ 380 ] who reported that phosphate supplementation during a 4-week weight loss program increased resting metabolic rate (RMR) and respiratory exchange ratio (suggesting greater carbohydrate utilization and caloric expenditure) during submaximal cycling exercise. In addition, Nazar and coworkers [ 381 ] reported that phosphate supplementation during an 8-week weight loss program increased RMR by 12-19% and prevented a normal decline in thyroid hormones. Although the rate of weight loss was similar in this trial, results suggest that phosphate supplementation may influence metabolic rate possibly by affecting thyroid hormones. Despite these to dated trials, no further research has been conducted and thus the role of phosphates in regards to weight loss is inconclusive at best.

Herbal Diuretics

This is a new type of supplement recently marketed as a natural way to promote weight loss. There is limited evidence that taraxacum officinale, verbena officinalis, lithospermum officinale, equisetum arvense, arctostaphylos uva-ursi, arctium lappa and silene saxifraga infusion may affect diuresis in animals [ 382 , 383 ]. Two studies presented at the 2001 American College of Sports Medicine meeting [ 384 , 385 ] indicated that although herbal diuretics promoted a small amount of dehydration (about 0.3% in one day), they were not nearly as effective as a common diuretic drug (about 3.1% dehydration in one day). Consequently, although more research is needed, the potential value of herbal diuretics as a weight loss supplement appears limited.

Performance Enhancement Supplements

A number of nutritional supplements have been proposed to enhance exercise performance. Some of these nutrients have been described above. Table 3 categorizes the proposed ergogenic nutrients into apparently safe and effective, possibly effective, too early to tell, and apparently ineffective. Weight gain supplements purported to increase muscle mass may also have ergogenic properties if they also promote increases in strength. Similarly, some sports may benefit from reductions in fat mass. Therefore, weight loss supplements that help athletes manage body weight and/or fat mass may also possess some ergogenic benefit. The following describes which supplements may or may not affect performance that were not previously described.

Water and Sports Drinks

Preventing dehydration during exercise is one of the keys of maintaining exercise performance (particularly in hot/humid environments). People engaged in intense exercise or work in the heat need to frequently ingest water or sports drinks (e.g., 1-2 cups every 10 - 15 minutes). The goal should be not to lose more than 2% of body weight during exercise (e.g., 180 lbs × 0.02 = 3.6 lbs). Sports drinks typically contain salt and carbohydrate at scientifically engendered quantities. Studies show that ingestion of sports drinks during exercise in hot/humid environments can help prevent dehydration and improve endurance exercise capacity [[ 56 ], von Duvillard 2005), [ 386 , 387 ]]. In fact, research has shown that carbohydrate intake during team sport type activities can increase exercise performance and CNS function [ 15 , 16 , 388 ]. Consequently, frequent ingestion of water and/or sports drinks during exercise is one of the easiest and most effective ergogenic aids.

One of the best ergogenic aids available for athletes and active individuals alike, is carbohydrate. Athletes and active individuals should consume a diet high in carbohydrate (e.g., 55 - 65% of calories or 5-8 grams/kg/day) in order to maintain muscle and liver carbohydrate stores [ 1 , 3 ]. Research has clearly identified carbohydrate is an ergogenic aid that can prolong exercise [ 3 ]. Additionally, ingesting a small amount of carbohydrate and protein 30-60 minutes prior to exercise and use of sports drinks during exercise can increase carbohydrate availability and improve exercise performance. Finally, ingesting carbohydrate and protein immediately following exercise can enhance carbohydrate storage and protein synthesis [ 1 , 3 ].

Earlier we indicated that creatine supplementation is one of the best supplements available to increase muscle mass and strength during training. However, creatine has also been reported to improve exercise capacity in a variety of events [[ 71 ], Kendall 2005, [ 389 – 391 ]]. This is particularly true when performing high intensity, intermittent exercise such as multiple sets of weight lifting, repeated sprints, and/or exercise involving sprinting and jogging (e.g., soccer) [ 71 ]. Creatine has also been shown to be effective at improving high intensity interval training. A 2009 study found that in addition to high intensity interval training creatine improved critical power [ 390 ]. Although studies evaluating the ergogenic value of creatine on endurance exercise perfor mance are mixed, endurance athletes may also theoretically benefit in several ways. For example, increasing creatine stores prior to carbohydrate loading (i.e., increasing dietary carbohydrate intake before competition in an attempt to maximize carbohydrate stores) has been shown to improve the ability to store carbohydrate [ 392 – 394 ]. A 2003 study found that ingesting 20 grams of creatine for 5 days improved endurance and anaerobic performance in elite rowers [ 395 ]. Further, co ingesting creatine with carbohydrate has been shown to optimize creatine and carbohydrate loading [ 396 ]. Most endurance athletes also perform interval training (sprint or speed work) in an attempt to improve anaerobic threshold. Since creatine has been reported to enhance interval sprint performance, creatine supplementation during training may improve training adaptations in endurance athletes [ 397 , 398 ]. Finally, many endurance athletes lose weight during their competitive season. Creatine supplementation during training may help people maintain weight.

Sodium Phosphate

We previously mentioned that sodium phosphate supplementation may increase resting energy expenditure and therefore could serve as a potential weight loss nutrient. However, most research on sodium phosphate has actually evaluated the potential ergogenic value. A number of studies indicated that sodium phosphate supplementation (e.g., 1 gram taken 4 times daily for 3-6 days) can increase maximal oxygen uptake (i.e., maximal aerobic capacity) and anaerobic threshold by 5-10% [ 399 – 403 ]. These finding suggest that sodium phosphate may be highly effective in improving endurance exercise capacity. In addition to endurance enhancement, sodium phosphate loading improved mean power output and oxygen uptake in trained cyclist in a 2008 study [ 404 ]. Other forms of phosphate (i.e., calcium phosphate, potassium phosphate) do not appear to possess ergogenic value.

Sodium Bicarbonate (Baking Soda)

During high intensity exercise, acid (H+) and carbon dioxide (CO 2 ) accumulate in the muscle and blood. One of the ways you get rid of the acidity and CO 2 is to buffer the acid and CO 2 with bicarbonate ions. The acid and CO 2 are then removed in the lungs. Bicarbonate loading (e.g., 0.3 grams per kg taken 60-90 minutes prior to exercise or 5 grams taken 2 times per day for 5-days) has been shown to be an effective way to buffer acidity during high intensity exercise lasting 1-3 minutes in duration [ 405 – 408 ]. This can improve exercise capacity in events like the 400 - 800 m run or 100 - 200 m swim [ 409 ]. In elite male swimmers sodium bicarbonate supplementation significantly improved 200 m freestyle performance [ 410 ]. A 2009 study found similar improvements in performance in youth swimmers at distances of 50 to 200 m. Although bicarbonate loading can improve exercise, some people have difficulty with their stomach tolerating bicarbonate as it may cause gastrointestinal distress.

Caffeine is a naturally derived stimulant found in many nutritional supplements typically as gaurana, bissey nut, or kola. Caffeine can also be found in coffee, tea, soft drinks, energy drinks, and chocolate. It has previously been made clear that caffeine can have a positive effect on energy expenditure, weight loss, and body fat. Caffeine has also been shown to be an effective ergogenic aid. Research investigating the effects of caffeine on a time trial in trained cyclist found that caffeine improved speed, peak power, and mean power [ 411 ]. Similar results were observed in a recent study that found cyclists who ingested a caffeine drink prior to a time trial demonstrated improvements in performance [ 412 , 413 ]. Studies indicate that ingestion of caffeine (e.g., 3-9 mg/kg taken 30 - 90 minutes before exercise) can spare carbohydrate use during exercise and thereby improve endurance exercise capacity [ 406 , 414 ]. In addition to the apparent positive effects on endurance performance, caffeine has also been shown to improve repeated sprint performance benefiting the anaerobic athlete [ 415 , 416 ]. People who drink caffeinated drinks regularly, however, appear to experience less ergogenic benefits from caffeine [ 417 ]. Additionally, some concern has been expressed that ingestion of caffeine prior to exercise may contribute to dehydration although recent studies have not supported this concern [ 414 , 418 , 419 ]. Caffeine doses above 9 mg/kg can result in urinary caffeine levels that surpass the doping threshold for many sport organizations. Suggestions that there is no ergogenic value to caffeine supplementation is not supported by the preponderance of available scientific studies.

In recent years research has begun investigating the effects of β-alanine supplementation on performance. β-alanine has ergogenic potential based on its relationship with carnosine. Carnosine is a dipeptide comprised of the amino acids, histidine and β-alanine naturally occurring in large amounts in skeletal muscles. Carnosine is believed to be one of the primary muscle-buffering substances available in skeletal muscle. Studies have demonstrated that taking β-alanine orally over a 28-day period was effective in increasing carnosine levels [ 420 , 421 ]. This proposed benefit would increase work capacity and decrease time to fatigue. Researchers have found that β-alanine supplementation decreases rate of fatigue [ 422 ]. This could translate into definite strength gains and improved performance. A recent study [ 423 ] supplemented men with β-alanine for 10 weeks and showed that muscle carnosine levels were significantly increased after 4 and 10 weeks of β-alanine supplementation.

Stout et al. [ 422 ] conducted a study that examined the effects of β-alanine supplementation on physical working capacity at fatigue threshold. The results showed decreased fatigue in the subjects tested. Other studies have shown that β-alanine supplementation can increase the number of repetitions one can do [ 424 ], increased lean body mass [ 425 ], increase knee extension torque [ 426 ] and training volume [ 427 ]. In fact, one study also showed that adding β-alanine supplementation with creatine improves performance over creatine alone [ 428 ]. While it appears that β-alanine supplementation can decrease fatigue rate, raise carnosine levels, and improve performance all of the research is not as favorable. There are other studies that show no performance benefits [ 425 , 429 ]

Post-Exercise Carbohydrate and Protein

Ingesting carbohydrate and protein following exercise enhances carbohydrate storage and protein synthesis. Theoretically, ingesting carbohydrate and protein following exercise may lead to greater training adaptations. In support of this theory, Esmarck and coworkers [ 107 ] found that ingesting carbohydrate and protein immediately following exercise doubled training adaptations in comparison to waiting until 2-hours to ingest carbohydrate and protein. Additionally, Tarnopolsky and associates [ 430 ] reported that post-exercise ingestion of carbohydrate with protein promoted as much strength gains as ingesting creatine with carbohydrate during training. A recent study by Kreider and colleagues [ 431 ] found that protein and carbohydrate supplementation post workout was capable of positively supporting the post exercise anabolic response. In the last few years many studies have agreed with these findings in that post workout supplementation is vital to recovery and training adaptations [ 13 , 104 , 431 – 433 ]. These findings underscore the importance of post-exercise carbohydrate and protein ingestion to support muscle anabolism and strength. However, it is still unclear if there are direct implications of protein/carbohydrate supplementation on other markers of performance such as time to exhaustion, maximal oxygen uptake, and/or skill development.

Ingestion of 3-6 grams of EAA following resistance exercise has been shown to increase protein synthesis [ 92 , 93 , 98 – 102 , 105 , 434 ]. Theoretically, ingestion of EAA after exercise should enhance gains in strength and muscle mass during training. While there is sound theoretical rationale, it is currently unclear whether following this strategy would lead to greater training adaptations and/or whether EAA supplementation would be better than simply ingesting carbohydrate and a quality protein following exercise.

Ingestion of BCAA (e.g., 6-10 grams per hour) with sports drinks during prolonged exercise would theoretically improve psychological perception of fatigue (i.e., central fatigue). Although there is strong rationale, the effects of BCAA supplementation on exercise performance is mixed with some studies suggesting an improvement and others showing no effect [ 33 ]. More research is needed before conclusions can be drawn.

HMB supplementation has been reported to improve training adaptations in untrained individuals initiating training as well as help reduce muscle breakdown in runners. Theoretically, this should enhance training adaptations in athletes. However, most studies show little benefit of HMB supplementation in athletes. A 2004 study by Hoffman [ 435 ] found HMB supplementation to be ineffective in collegiate football players after short term supplementation. It has been hypothesized that HMB will delay or prevent muscle damage; however this has limited evidence as suggested in previous sections. There are a few studies that have been positive [ 115 ]. A 2009 study found that HMB supplementation did positively affect strength in trained men [ 436 ]. While HMB supplementation may still have some scientific rationale there is little evidence that is can directly affect performance in moderately trained subjects.

Ingesting glycerol with water has been reported to increase fluid retention [ 437 ]. Theoretically, this should help athletes prevent dehydration during prolonged exercise and improve performance particularly if they are susceptible to dehydration. Although studies indicate that glycerol can significantly enhance body fluid, results are mixed on whether it can improve exercise capacity [ 69 , 438 – 443 ]. Little research has been done on glycerol in the last five years however, a 2006 study agreed with previous findings in that glycerol has little impact on performance [ 444 ].

A number of supplements purported to enhance performance and/or training adaptation fall under this category. This includes the weight gain and weight loss supplements listed in Table 3 as well as the following supplements not previously described in this category.

Medium Chain Triglycerides (MCT)

MCT's are shorter chain fatty acids that can easily enter the mitochondria of the cell and be converted to energy through fat metabolism [ 445 ]. Studies are mixed as to whether MCT's can serve as an effective source of fat during exercise metabolism and/or improve exercise performance [ 445 – 449 ]. A 2001 study found that 60 g/day of MCT oil for two weeks was not sufficient at improving performance [ 450 ]. In fact Goedecke found that not only did MCT supplementation not improve performance, but, actually negatively affected sprint performance in trained cyclists [ 451 ]. These findings have been confirmed by others that MCT oils are not sufficient to induce positive training adaptations and may cause gastric distress [ 452 , 453 ]. It must be noted that while most studies have not been favourable, one 2009 study found that MCT oil may positively affect RPE and lactate clearance [ 454 ]. It does not appear likely that MCT can positively affect training adaptations, but further research is needed.

As described above, glutamine has been shown to influence protein synthesis and help maintain the immune system. Theoretically, glutamine supplementation during training should enhance gains in strength and muscle mass as well as help athletes tolerate training to a better degree. Although there is some evidence that glutamine supplementation with protein can improve training adaptations, more research is needed to determine the ergogenic value in athletes. There is currently no research to suggest that glutamine has a direct effect on performance.

Ribose is a 3-carbon carbohydrate that is involved in the synthesis of adenosine triphosphate (ATP) in the muscle (the useable form of energy). Clinical studies have shown that ribose supplementation can increase exercise capacity in heart patients [ 455 – 459 ]. For this reason, ribose has been suggested to be an ergogenic aid for athletes. Although more research is needed, most studies show no ergogenic value of ribose supplementation on exercise capacity in health untrained or trained populations [ 460 – 462 ]. A 2006 study [ 463 ] investigated the effects of ribose vs. dextrose on rowing performance. After eight weeks of supplementation dextrose had a better response than ribose across the subjects [ 463 ]. Kreider and associates [ 462 ] and Kersick and colleagues [ 464 ] investigated ribose supplementation on measures of anaerobic capacity in trained athletes. This research group found that ribose supplementation did not have a positive impact on performance [ 462 , 464 ]. It appears at this point that ribose supplementation does not improve aerobic or anaerobic performance.

Inosine is a building block for DNA and RNA that is found in muscle. Inosine has a number of potentially important roles that may enhance training and/or exercise performance [ 465 ]. Although there is some theoretical rationale, available studies indicate that inosine supplementation has no apparent affect on exercise performance capacity [ 466 – 468 ].

Supplements to Promote General Health

In addition to the supplements previously described, several nutrients have been suggested to help athletes stay healthy during intense training. For example, the American Medical Association recently recommended that all Americans ingest a daily low-dose multivitamin in order to ensure that people get a sufficient amount of vitamins and minerals in their diet. Although one-a-day vitamin supplementation has not been found to improve exercise capacity in athletes, it may make sense to take a daily vitamin supplement for health reasons. Glucosomine and chondroitin have been reported to slow cartilage degeneration and reduce the degree of joint pain in active individuals which may help athletes postpone and/or prevent joint problems [ 469 , 470 ]. Supplemental Vitamin C, glutamine, echinacea, and zinc have been reported to enhance immune function [ 471 – 474 ]. Consequently, some sports nutritionists recommend that athletes who feel a cold coming on take these nutrients in order to enhance immune function [ 55 , 471 – 473 ]. Similarly, although additional research is necessary, Vitamin E, Vitamin C, selenium, alpha-lipoic acid and other antioxidants may help restore overwhelmed anti-oxidant defences exhibited by athletes and reduce the risk of numerous chronic diseases in some instances [ 475 ]. Creatine, calcium β-HMB, BCAA, and L-carnitine tartrate have been shown to help athletes tolerate heavy training periods [ 31 , 118 , 125 , 126 , 128 , 379 , 476 – 478 ]. Finally, the omega-3 fatty acids docosahexaenoic acid (DHA) and eicosapantaenoic acid (EPA), in supplemental form, are now endorsed by the American Heart Association for heart health in certain individuals [ 479 ]. This supportive supplement position stems from: 1.) an inability to consume cardio-protective amounts by diet alone; and, 2.) the mercury contamination sometimes present in whole-food sources of DHA and EPA found in fatty fish. Consequently, prudent use of these types of nutrients at various times during training may help athletes stay healthy and/or tolerate training to a greater degree [ 50 ].

Maintaining an energy balance and nutrient dense diet, prudent training, proper timing of nutrient intake, and obtaining adequate rest are the cornerstones to enhancing performance and/or training adaptations. Use of a limited number of nutritional supplements that research has supported can help improve energy availability (e.g., sports drinks, carbohydrate, creatine, caffeine, β-alanine, etc) and/or promote recovery (carbohydrate, protein, essential amino acids, etc) can provide additional benefit in certain instances. The sports nutrition specialist should stay up to date regarding the role of nutrition on exercise so they can provide honest and accurate information to their students, clients, and/or athletes about the role of nutrition and dietary supplements on performance and training. Furthermore, the sports nutrition specialist should actively participate in exercise nutrition research; write unbiased scholarly reviews for journals and lay publications; help disseminate the latest research findings to the public so they can make informed decisions about appropriate methods of exercise, dieting, and/or whether various nutritional supplements can affect health, performance, and/or training; and, disclose any commercial or financial conflicts of interest during such promulgations to the public. Finally, companies selling nutritional supplements should develop scientifically based products, conduct research on their products, and honestly market the results of studies so consumers can make informed decisions.

Leutholtz B, Kreider R: Exercise and Sport Nutrition. Nutritional Health. Edited by: Wilson T, Temple N. 2001, Totowa, NJ: Humana Press, 207-39.

Chapter   Google Scholar  

Williams MH: Nutrition for Health, Fitness, and Sport. 1999, Dubuque, IA: ACB/McGraw-Hill

Google Scholar  

Kreider R, Leutholtz B, Katch F, Katch V: Exercise & Sport Nutrition. 2009, Santa Barbara: Fitness Technologies Press

FDA: Dietary Supplements. 2003, [ http://www.cfsan.fda.gov/~dms/ds-faq.html ]

Beers MH, Berkow R: The Merck Manual. 1999, Merck Research Laboratories, 17

Sherman WM, Jacobs KA, Leenders N: Carbohydrate metabolism during endurance exercise. Overtraining in Sport. Edited by: Kreider RB, Fry AC, O'Toole ML. 1998, Champaign: Human Kinetics Publishers, 289-308.

Berning JR: Energy intake, diet, and muscle wasting. Overtraining in Sport. Edited by: Kreider RB, Fry AC, O'Toole ML. 1998, Champaign: Human Kinetics, 275-88.

Kreider RB, Fry AC, O'Toole ML: Overtraining in Sport. 1998, Champaign: Human Kinetics Publishers

Kreider RB: Physiological considerations of ultraendurance performance. Int J Sport Nutr. 1991, 1 (1): 3-27.

CAS   PubMed   Google Scholar  

Brouns F, Saris WH, Beckers E, Adlercreutz H, Vusse van der GJ, Keizer HA, Kuipers H, Menheere P, Wagenmakers AJ, ten Hoor F: Metabolic changes induced by sustained exhaustive cycling and diet manipulation. Int J Sports Med. 1989, 10 (Suppl 1): S49-62. 10.1055/s-2007-1024954.

Article   PubMed   Google Scholar  

Brouns F, Saris WH, Stroecken J, Beckers E, Thijssen R, Rehrer NJ, ten Hoor F: Eating, drinking, and cycling. A controlled Tour de France simulation study, Part I. Int J Sports Med. 1989, 10 (Suppl 1): S32-40. 10.1055/s-2007-1024952.

Brouns F, Saris WH, Stroecken J, Beckers E, Thijssen R, Rehrer NJ, ten Hoor F: Eating, drinking, and cycling. A controlled Tour de France simulation study Part II. Effect of diet manipulation. Int J Sports Med. 1989, 10 (Suppl 1): S41-8. 10.1055/s-2007-1024953.

Kerksick C, Harvey T, Stout J, Campbell B, Wilborn C, Kreider R, Kalman D, Ziegenfuss T, Lopez H, Landis J, Ivy JL, Antonio J: International Society of Sports Nutrition position stand: nutrient timing. J Int Soc Sports Nutr. 2008, 5: 17-10.1186/1550-2783-5-17.

Article   PubMed Central   PubMed   CAS   Google Scholar  

Harger-Domitrovich SG, McClaughry AE, Gaskill SE, Ruby BC: Exogenous carbohydrate spares muscle glycogen in men and women during 10 h of exercise. Med Sci Sports Exerc. 2007, 39 (12): 2171-9. 10.1249/mss.0b013e318157a650.

Article   CAS   PubMed   Google Scholar  

Rodriguez NR, Di Marco NM, Langley S: American College of Sports Medicine position stand. Nutrition and athletic performance. Med Sci Sports Exerc. 2009, 41 (3): 709-31. 10.1249/MSS.0b013e31890eb86.

Article   PubMed   CAS   Google Scholar  

Rodriguez NR, DiMarco NM, Langley S: Position of the American Dietetic Association, Dietitians of Canada, and the American College of Sports Medicine: Nutrition and athletic performance. J Am Diet Assoc. 2009, 109 (3): 509-27. 10.1016/j.jada.2009.01.005.

Sawka MN, Burke LM, Eichner ER, Maughan RJ, Montain SJ, Stachenfeld NS: American College of Sports Medicine position stand. Exercise and fluid replacement. Med Sci Sports Exerc. 2007, 39 (2): 377-90. 10.1249/mss.0b013e31802ca597.

Currell K, Jeukendrup AE: Superior endurance performance with ingestion of multiple transportable carbohydrates. Med Sci Sports Exerc. 2008, 40 (2): 275-81. 10.1249/mss.0b013e31815adf19.

Jeukendrup AE, Moseley L: Multiple transportable carbohydrates enhance gastric emptying and fluid delivery. Scand J Med Sci Sports. 2008

Earnest CP, Lancaster SL, Rasmussen CJ, Kerksick CM, Lucia A, Greenwood MC, Almada AL, Cowan PA, Kreider RB: Low vs. high glycemic index carbohydrate gel ingestion during simulated 64-km cycling time trial performance. J Strength Cond Res. 2004, 18 (3): 466-72. 10.1519/R-xxxxx.1.

PubMed   Google Scholar  

Venables MC, Brouns F, Jeukendrup AE: Oxidation of maltose and trehalose during prolonged moderate-intensity exercise. Med Sci Sports Exerc. 2008, 40 (9): 1653-9. 10.1249/MSS.0b013e318175716c.

Jentjens RL, Jeukendrup AE: Effects of pre-exercise ingestion of trehalose, galactose and glucose on subsequent metabolism and cycling performance. Eur J Appl Physiol. 2003, 88 (4-5): 459-65. 10.1007/s00421-002-0729-7.

Achten J, Jentjens RL, Brouns F, Jeukendrup AE: Exogenous oxidation of isomaltulose is lower than that of sucrose during exercise in men. J Nutr. 2007, 137 (5): 1143-8.

Jentjens RL, Venables MC, Jeukendrup AE: Oxidation of exogenous glucose, sucrose, and maltose during prolonged cycling exercise. J Appl Physiol. 2004, 96 (4): 1285-91. 10.1152/japplphysiol.01023.2003.

Jeukendrup AE, Jentjens R: Oxidation of carbohydrate feedings during prolonged exercise: current thoughts, guidelines and directions for future research. Sports Med. 2000, 29 (6): 407-24. 10.2165/00007256-200029060-00004.

Rowlands DS, Wallis GA, Shaw C, Jentjens RL, Jeukendrup AE: Glucose polymer molecular weight does not affect exogenous carbohydrate oxidation. Med Sci Sports Exerc. 2005, 37 (9): 1510-6. 10.1249/01.mss.0000177586.68399.f5.

Lemon PW, Tarnopolsky MA, MacDougall JD, Atkinson SA: Protein requirements and muscle mass/strength changes during intensive training in novice bodybuilders. J Appl Physiol. 1992, 73 (2): 767-75.

Tarnopolsky MA, MacDougall JD, Atkinson SA: Influence of protein intake and training status on nitrogen balance and lean body mass. J Appl Physiol. 1988, 64 (1): 187-93.

Tarnopolsky MA, Atkinson SA, MacDougall JD, Chesley A, Phillips S, Schwarcz HP: Evaluation of protein requirements for trained strength athletes. J Appl Physiol. 1992, 73 (5): 1986-95.

Tarnopolsky MA: Protein and physical performance. Curr Opin Clin Nutr Metab Care. 1999, 2 (6): 533-7. 10.1097/00075197-199911000-00018.

Kreider RB: Dietary supplements and the promotion of muscle growth with resistance exercise. Sports Med. 1999, 27 (2): 97-110. 10.2165/00007256-199927020-00003.

Chesley A, MacDougall JD, Tarnopolsky MA, Atkinson SA, Smith K: Changes in human muscle protein synthesis after resistance exercise. J Appl Physiol. 1992, 73 (4): 1383-8.

Kreider RB: Effects of protein and amino acid supplementation on athletic performance. Sportscience. 1999, 3 (1): [ http://www.sportsci.org/jour/9901/rbk.html ]

Kreider RB, Kleiner SM: Protein supplements for athletes: need vs. convenience. Your Patient & Fitness. 2000, 14 (6): 12-8.

Bucci L, Unlu L: Proteins and amino acid supplements in exercise and sport. Energy-Yielding Macronutrients and Energy Metabolism in Sports Nutrition. Edited by: Driskell J, Wolinsky I. 2000, Boca Raton, FL: CRC Press, 191-212.

Boirie Y, Dangin M, Gachon P, Vasson MP, Maubois JL, Beaufrere B: Slow and fast dietary proteins differently modulate postprandial protein accretion. Proc Natl Acad Sci USA. 1997, 94 (26): 14930-5. 10.1073/pnas.94.26.14930.

Article   PubMed Central   CAS   PubMed   Google Scholar  

Boirie Y, Beaufrere B, Ritz P: Energetic cost of protein turnover in healthy elderly humans. Int J Obes Relat Metab Disord. 2001, 25 (5): 601-5. 10.1038/sj.ijo.0801608.

Boirie Y, Gachon P, Corny S, Fauquant J, Maubois JL, Beaufrere B: Acute postprandial changes in leucine metabolism as assessed with an intrinsically labeled milk protein. Am J Physiol. 1996, 271 (6 Pt 1): E1083-91.

Campbell B, Kreider RB, Ziegenfuss T, La Bounty P, Roberts M, Burke D, Landis J, Lopez H, Antonio J: International Society of Sports Nutrition position stand: protein and exercise. J Int Soc Sports Nutr. 2007, 4: 8-10.1186/1550-2783-4-8.

Article   PubMed Central   PubMed   Google Scholar  

Venkatraman JT, Leddy J, Pendergast D: Dietary fats and immune status in athletes: clinical implications. Med Sci Sports Exerc. 2000, 32 (7 Suppl): S389-95.

Dorgan JF, Judd JT, Longcope C, Brown C, Schatzkin A, Clevidence BA, Campbell WS, Nair PP, Franz C, Kahle L, Taylor PR: Effects of dietary fat and fiber on plasma and urine androgens and estrogens in men: a controlled feeding study. Am J Clin Nutr. 1996, 64 (6): 850-5.

Hamalainen EK, Adlercreutz H, Puska P, Pietinen P: Decrease of serum total and free testosterone during a low-fat high-fibre diet. J Steroid Biochem. 1983, 18 (3): 369-70. 10.1016/0022-4731(83)90117-6.

Reed MJ, Cheng RW, Simmonds M, Richmond W, James VH: Dietary lipids: an additional regulator of plasma levels of sex hormone binding globulin. J Clin Endocrinol Metab. 1987, 64 (5): 1083-5. 10.1210/jcem-64-5-1083.

Fry AC, Kraemer WJ, Ramsey LT: Pituitary-adrenal-gonadal responses to high-intensity resistance exercise overtraining. J Appl Physiol. 1998, 85 (6): 2352-9.

Miller WC, Koceja DM, Hamilton EJ: A meta-analysis of the past 25 years of weight loss research using diet, exercise or diet plus exercise intervention. Int J Obes Relat Metab Disord. 1997, 21 (10): 941-7. 10.1038/sj.ijo.0800499.

Miller WC: Effective diet and exercise treatments for overweight and recommendations for intervention. Sports Med. 2001, 31 (10): 717-24. 10.2165/00007256-200131100-00002.

Pirozzo S, Summerbell C, Cameron C, Glasziou P: Should we recommend low-fat diets for obesity?. Obes Rev. 2003, 4 (2): 83-90. 10.1046/j.1467-789X.2003.00099.x.

Hu FB, Manson JE, Willett WC: Types of dietary fat and risk of coronary heart disease: a critical review. J Am Coll Nutr. 2001, 20 (1): 5-19.

Vessby B: Dietary fat, fatty acid composition in plasma and the metabolic syndrome. Curr Opin Lipidol. 2003, 14 (1): 15-9. 10.1097/00041433-200302000-00004.

Kreider RB: Effects of creatine supplementation on performance and training adaptations. Abstracts of 6th Internationl Conference on Guanidino Compounds in Biology and Medicine. 2001

Carli G, Bonifazi M, Lodi L, Lupo C, Martelli G, Viti A: Changes in the exercise-induced hormone response to branched chain amino acid administration. Eur J Appl Physiol Occup Physiol. 1992, 64 (3): 272-7. 10.1007/BF00626291.

Cade JR, Reese RH, Privette RM, Hommen NM, Rogers JL, Fregly MJ: Dietary intervention and training in swimmers. Eur J Appl Physiol Occup Physiol. 1991, 63 (3-4): 210-5. 10.1007/BF00233850.

Nieman DC, Fagoaga OR, Butterworth DE, Warren BJ, Utter A, Davis JM, Henson DA, Nehlsen-Cannarella SL: Carbohydrate supplementation affects blood granulocyte and monocyte trafficking but not function after 2.5 h or running. Am J Clin Nutr. 1997, 66 (1): 153-9.

Nieman DC: Influence of carbohydrate on the immune response to intensive, prolonged exercise. Exerc Immunol Rev. 1998, 4: 64-76.

Nieman DC: Nutrition, exercise, and immune system function. Clin Sports Med. 1999, 18 (3): 537-48. 10.1016/S0278-5919(05)70167-8.

Burke LM: Nutritional needs for exercise in the heat. Comp Biochem Physiol A Mol Integr Physiol. 2001, 128 (4): 735-48. 10.1016/S1095-6433(01)00279-3.

Burke LM: Nutrition for post-exercise recovery. Aust J Sci Med Sport. 1997, 29 (1): 3-10.

Maughan RJ, Noakes TD: Fluid replacement and exercise stress. A brief review of studies on fluid replacement and some guidelines for the athlete. Sports Med. 1991, 12 (1): 16-31. 10.2165/00007256-199112010-00003.

Zawadzki KM, Yaspelkis BB, Ivy JL: Carbohydrate-protein complex increases the rate of muscle glycogen storage after exercise. J Appl Physiol. 1992, 72 (5): 1854-9.

Tarnopolsky MA, Bosman M, Macdonald JR, Vandeputte D, Martin J, Roy BD: Postexercise protein-carbohydrate and carbohydrate supplements increase muscle glycogen in men and women. J Appl Physiol. 1997, 83 (6): 1877-83.

Kraemer WJ, Volek JS, Bush JA, Putukian M, Sebastianelli WJ: Hormonal responses to consecutive days of heavy-resistance exercise with or without nutritional supplementation. J Appl Physiol. 1998, 85 (4): 1544-55.

Jeukendrup AE, Currell K, Clarke J, Cole J, Blannin AK: Effect of beverage glucose and sodium content on fluid delivery. Nutr Metab (Lond). 2009, 6: 9-10.1186/1743-7075-6-9.

Article   CAS   Google Scholar  

Rehrer NJ: Fluid and electrolyte balance in ultra-endurance sport. Sports Med. 2001, 31 (10): 701-15. 10.2165/00007256-200131100-00001.

Sawka MN, Montain SJ: Fluid and electrolyte supplementation for exercise heat stress. Am J Clin Nutr. 2000, 72 (2 Suppl): 564S-72S.

Shirreffs SM, Armstrong LE, Cheuvront SN: Fluid and electrolyte needs for preparation and recovery from training and competition. J Sports Sci. 2004, 22 (1): 57-63. 10.1080/0264041031000140572.

Brouns F, Kovacs EM, Senden JM: The effect of different rehydration drinks on post-exercise electrolyte excretion in trained athletes. Int J Sports Med. 1998, 19 (1): 56-60. 10.1055/s-2007-971881.

Kovacs EM, Senden JM, Brouns F: Urine color, osmolality and specific electrical conductance are not accurate measures of hydration status during postexercise rehydration. J Sports Med Phys Fitness. 1999, 39 (1): 47-53.

Kovacs EM, Schmahl RM, Senden JM, Brouns F: Effect of high and low rates of fluid intake on post-exercise rehydration. Int J Sport Nutr Exerc Metab. 2002, 12 (1): 14-23.

Meyer LG, Horrigan DJ, Lotz WG: Effects of three hydration beverages on exercise performance during 60 hours of heat exposure. Aviat Space Environ Med. 1995, 66 (11): 1052-7.

Williams MH: Facts and fallacies of purported ergogenic amino acid supplements. Clin Sports Med. 1999, 18 (3): 633-49. 10.1016/S0278-5919(05)70173-3.

Kreider RB: Effects of creatine supplementation on performance and training adaptations. Mol Cell Biochem. 2003, 244 (1-2): 89-94. 10.1023/A:1022465203458.

Volek JS, Duncan ND, Mazzetti SA, Putukian M, Gomez AL, Staron RS, Kraemer WJ: Performance and muscle fiber adaptations to 12 weeks of creatine supplementation and heavy resistance training. Medicine & Science in Sports & Exercise. 1999, 31 (5):

Willoughby DS, Rosene J: Effects of oral creatine and resistance training on myosin heavy chain expression. Med Sci Sports Exerc. 2001, 33 (10): 1674-81. 10.1097/00005768-200110000-00010.

Willoughby DS, Rosene JM: Effects of oral creatine and resistance training on myogenic regulatory factor expression. Med Sci Sports Exerc. 2003, 35 (6): 923-9. 10.1249/01.MSS.0000069746.05241.F0.

Olsen S, Aagaard P, Kadi F, Tufekovic G, Verney J, Olesen JL, Suetta C, Kjaer M: Creatine supplementation augments the increase in satellite cell and myonuclei number in human skeletal muscle induced by strength training. J Physiol. 2006, 573 (Pt 2): 525-34. 10.1113/jphysiol.2006.107359.

Williams MH, Kreider R, Branch JD: Creatine: The power supplement. 1999, Champaign, IL: Human Kinetics Publishers

Kreider R, Melton C, Hunt J, Rasmussen C, Ransom J, Stroud T, Cantler E, Milnor P: Creatine does not increase incidence of cramping or injury during pre-season college football training I. Med Sci Sports Exerc. 1999, 31 (5): S355-

Article   Google Scholar  

Kreider RB, Melton C, Rasmussen CJ, Greenwood M, Lancaster S, Cantler EC, Milnor P, Almada AL: Long-term creatine supplementation does not significantly affect clinical markers of health in athletes. Mol Cell Biochem. 2003, 244 (1-2): 95-104. 10.1023/A:1022469320296.

Graham AS, Hatton RC: Creatine: a review of efficacy and safety. J Am Pharm Assoc (Wash). 1999, 39 (6): 803-10.

CAS   Google Scholar  

Juhn MS, Tarnopolsky M: Potential side effects of oral creatine supplementation: a critical review. Clin J Sport Med. 1998, 8 (4): 298-304.

Taes YE, Delanghe JR, Wuyts B, Voorde Van De J, Lameire NH: Creatine supplementation does not affect kidney function in an animal model with pre-existing renal failure. Nephrol Dial Transplant. 2003, 18 (2): 258-64. 10.1093/ndt/18.2.258.

Schilling BK, Stone MH, Utter A, Kearney JT, Johnson M, Coglianese R, Smith L, O'Bryant HS, Fry AC, Starks M, Keith R, Stone ME: Creatine supplementation and health variables: a retrospective study. Med Sci Sports Exerc. 2001, 33 (2): 183-8.

Greenwood M, Kreider R, Greenwood L, Byars A: Creatine supplementation does not increase the incidence of injury or cramping in college baseball players. Journal of Exercise Physiology online. 2003, 6 (4): 16-22.

Greenwood M, Kreider R, Greenwood L, Earnest C, Farris J, Brown L: Effects of creatine supplementation on the incidence of cramping/injury during eighteen weeks of collegiate baseball training/competition. Med Sci Sport Exerc. 2002, 34 (S146):

Watsford ML, Murphy AJ, Spinks WL, Walshe AD: Creatine supplementation and its effect on musculotendinous stiffness and performance. J Strength Cond Res. 2003, 17 (1): 26-33. 10.1519/1533-4287(2003)017<0026:CSAIEO>2.0.CO;2.

Dalbo VJ, Roberts MD, Stout JR, Kerksick CM: Putting to rest the myth of creatine supplementation leading to muscle cramps and dehydration. Br J Sports Med. 2008, 42 (7): 567-73. 10.1136/bjsm.2007.042473.

Buford TW, Kreider RB, Stout JR, Greenwood M, Campbell B, Spano M, Ziegenfuss T, Lopez H, Landis J, Antonio J: International Society of Sports Nutrition position stand: creatine supplementation and exercise. J Int Soc Sports Nutr. 2007, 4: 6-10.1186/1550-2783-4-6.

Brown EC, DiSilvestro RA, Babaknia A, Devor ST: Soy versus whey protein bars: effects on exercise training impact on lean body mass and antioxidant status. Nutr J. 2004, 3: 22-10.1186/1475-2891-3-22.

Candow DG, Burke NC, Smith-Palmer T, Burke DG: Effect of whey and soy protein supplementation combined with resistance training in young adults. Int J Sport Nutr Exerc Metab. 2006, 16 (3): 233-44.

Flakoll PJ, Judy T, Flinn K, Carr C, Flinn S: Postexercise protein supplementation improves health and muscle soreness during basic military training in Marine recruits. J Appl Physiol. 2004, 96 (3): 951-6. 10.1152/japplphysiol.00811.2003.

Kalman D, Feldman S, Martinez M, Krieger DR, Tallon MJ: Effect of protein source and resistance training on body composition and sex hormones. J Int Soc Sports Nutr. 2007, 4: 4-10.1186/1550-2783-4-4.

Biolo G, Williams BD, Fleming RY, Wolfe RR: Insulin action on muscle protein kinetics and amino acid transport during recovery after resistance exercise. Diabetes. 1999, 48 (5): 949-57. 10.2337/diabetes.48.5.949.

Borsheim E, Tipton KD, Wolf SE, Wolfe RR: Essential amino acids and muscle protein recovery from resistance exercise. Am J Physiol Endocrinol Metab. 2002, 283 (4): E648-57.

Burk A, Timpmann S, Medijainen L, Vahi M, Oopik V: Time-divided ingestion pattern of casein-based protein supplement stimulates an increase in fat-free body mass during resistance training in young untrained men. Nutr Res. 2009, 29 (6): 405-13. 10.1016/j.nutres.2009.03.008.

Cribb PJ, Williams AD, Carey MF, Hayes A: The effect of whey isolate and resistance training on strength, body composition, and plasma glutamine. Int J Sport Nutr Exerc Metab. 2006, 16 (5): 494-509.

Hoffman JR, Ratamess NA, Tranchina CP, Rashti SL, Kang J, Faigenbaum AD: Effect of protein-supplement timing on strength, power, and body-composition changes in resistance-trained men. Int J Sport Nutr Exerc Metab. 2009, 19 (2): 172-85.

Holm L, Olesen JL, Matsumoto K, Doi T, Mizuno M, Alsted TJ, Mackey AL, Schwarz P, Kjaer M: Protein-containing nutrient supplementation following strength training enhances the effect on muscle mass, strength, and bone formation in postmenopausal women. J Appl Physiol. 2008, 105 (1): 274-81. 10.1152/japplphysiol.00935.2007.

Kobayashi H, Borsheim E, Anthony TG, Traber DL, Badalamenti J, Kimball SR, Jefferson LS, Wolfe RR: Reduced amino acid availability inhibits muscle protein synthesis and decreases activity of initiation factor eIF2B. Am J Physiol Endocrinol Metab. 2003, 284 (3): E488-98.

Miller SL, Tipton KD, Chinkes DL, Wolf SE, Wolfe RR: Independent and combined effects of amino acids and glucose after resistance exercise. Med Sci Sports Exerc. 2003, 35 (3): 449-55. 10.1249/01.MSS.0000053910.63105.45.

Rasmussen BB, Tipton KD, Miller SL, Wolf SE, Wolfe RR: An oral essential amino acid-carbohydrate supplement enhances muscle protein anabolism after resistance exercise. J Appl Physiol. 2000, 88 (2): 386-92.

Rasmussen BB, Wolfe RR, Volpi E: Oral and intravenously administered amino acids produce similar effects on muscle protein synthesis in the elderly. J Nutr Health Aging. 2002, 6 (6): 358-62.

PubMed Central   CAS   PubMed   Google Scholar  

Tipton KD, Rasmussen BB, Miller SL, Wolf SE, Owens-Stovall SK, Petrini BE, Wolfe RR: Timing of amino acid-carbohydrate ingestion alters anabolic response of muscle to resistance exercise. Am J Physiol Endocrinol Metab. 2001, 281 (2): E197-206.

Verdijk LB, Jonkers RA, Gleeson BG, Beelen M, Meijer K, Savelberg HH, Wodzig WK, Dendale P, van Loon LJ: Protein supplementation before and after exercise does not further augment skeletal muscle hypertrophy after resistance training in elderly men. Am J Clin Nutr. 2009, 89 (2): 608-16. 10.3945/ajcn.2008.26626.

Willoughby DS, Stout JR, Wilborn CD: Effects of resistance training and protein plus amino acid supplementation on muscle anabolism, mass, and strength. Amino Acids. 2007, 32 (4): 467-77. 10.1007/s00726-006-0398-7.

Wolfe RR: Regulation of muscle protein by amino acids. J Nutr. 2002, 132 (10): 3219S-24S.

Tipton KD, Borsheim E, Wolf SE, Sanford AP, Wolfe RR: Acute response of net muscle protein balance reflects 24-h balance after exercise and amino acid ingestion. Am J Physiol Endocrinol Metab. 2003, 284 (1): E76-89.

Esmarck B, Andersen JL, Olsen S, Richter EA, Mizuno M, Kjaer M: Timing of postexercise protein intake is important for muscle hypertrophy with resistance training in elderly humans. J Physiol. 2001, 535 (Pt 1): 301-11. 10.1111/j.1469-7793.2001.00301.x.

Garlick PJ: The role of leucine in the regulation of protein metabolism. J Nutr. 2005, 135 (6 Suppl): 1553S-6S.

Garlick PJ, Grant I: Amino acid infusion increases the sensitivity of muscle protein synthesis in vivo to insulin. Effect of branched-chain amino acids. Biochem J. 1988, 254 (2): 579-84.

Nair KS: Leucine as a regulator of whole body and skeletal muscle protein metabolism in humans. Am J Physiol. 1992, 263 (5 Pt 1): E928-34.

Wilson GJ, Wilson JM, Manninen AH: Effects of beta-hydroxy-beta-methylbutyrate (HMB) on exercise performance and body composition across varying levels of age, sex, and training experience: A review. Nutr Metab (Lond). 2008, 5: 1. 10.1186/1743-7075-5-1.

Gallagher PM, Carrithers JA, Godard MP, Schulze KE, Trappe SW: Beta-hydroxy-beta-methylbutyrate ingestion, Part I: effects on strength and fat free mass. Med Sci Sports Exerc. 2000, 32 (12): 2109-15. 10.1097/00005768-200012000-00022.

Gallagher PM, Carrithers JA, Godard MP, Schulze KE, Trappe SW: Beta-hydroxy-beta-methylbutyrate ingestion, part II: effects on hematology, hepatic and renal function. Med Sci Sports Exerc. 2000, 32 (12): 2116-9. 10.1097/00005768-200012000-00023.

Nissen S, Sharp R, Ray M: Effect of leucine metabolite beta-hydroxy-beta-methylbutyrate on muscle metabolism during resistance exercise testing. J Am Physiol. 1996, 81: 2095-104.

Panton LB, Rathmacher JA, Baier S, Nissen S: Nutritional supplementation of the leucine metabolite beta-hydroxy-beta-methylbutyrate (hmb) during resistance training. Nutrition. 2000, 16 (9): 734-9. 10.1016/S0899-9007(00)00376-2.

Slater GJ, Jenkins D: Beta-hydroxy-beta-methylbutyrate (HMB) supplementation and the promotion of muscle growth and strength. Sports Med. 2000, 30 (2): 105-16. 10.2165/00007256-200030020-00004.

Vukovich MD, Stubbs NB, Bohlken RM: Body composition in 70-year-old adults responds to dietary beta-hydroxy-beta-methylbutyrate similarly to that of young adults. J Nutr. 2001, 131 (7): 2049-52.

Knitter AE, Panton L, Rathmacher JA, Petersen A, Sharp R: Effects of beta-hydroxy-beta-methylbutyrate on muscle damage after a prolonged run. J Appl Physiol. 2000, 89 (4): 1340-4.

Smith HJ, Wyke SM, Tisdale MJ: Mechanism of the attenuation of proteolysis-inducing factor stimulated protein degradation in muscle by beta-hydroxy-beta-methylbutyrate. Cancer Res. 2004, 64 (23): 8731-5. 10.1158/0008-5472.CAN-04-1760.

Jowko E, Ostaszewski P, Jank M, Sacharuk J, Zieniewicz A, Wilczak J, Nissen S: Creatine and beta-hydroxy-beta-methylbutyrate (HMB) additively increase lean body mass and muscle strength during a weight-training program. Nutrition. 2001, 17 (7-8): 558-66. 10.1016/S0899-9007(01)00540-8.

O'Connor DM, Crowe MJ: Effects of beta-hydroxy-beta-methylbutyrate and creatine monohydrate supplementation on the aerobic and anaerobic capacity of highly trained athletes. J Sports Med Phys Fitness. 2003, 43 (1): 64-8.

Kreider RB, Ferreira M, Wilson M, Almada AL: Effects of calcium beta-hydroxy-beta-methylbutyrate (HMB) supplementation during resistance-training on markers of catabolism, body composition and strength. Int J Sports Med. 1999, 20 (8): 503-9. 10.1055/s-1999-8835.

Slater G, Jenkins D, Logan P, Lee H, Vukovich M, Rathmacher JA, Hahn AG: Beta-hydroxy-beta-methylbutyrate (HMB) supplementation does not affect changes in strength or body composition during resistance training in trained men. Int J Sport Nutr Exerc Metab. 2001, 11 (3): 384-96.

Ransone J, Neighbors K, Lefavi R, Chromiak J: The effect of beta-hydroxy beta-methylbutyrate on muscular strength and body composition in collegiate football players. J Strength Cond Res. 2003, 17 (1): 34-9. 10.1519/1533-4287(2003)017<0034:TEOHMO>2.0.CO;2.

Coombes JS, McNaughton LR: Effects of branched-chain amino acid supplementation on serum creatine kinase and lactate dehydrogenase after prolonged exercise. J Sports Med Phys Fitness. 2000, 40 (3): 240-6.

Schena F, Guerrini F, Tregnaghi P, Kayser B: Branched-chain amino acid supplementation during trekking at high altitude. The effects on loss of body mass, body composition, and muscle power. Eur J Appl Physiol Occup Physiol. 1992, 65 (5): 394-8. 10.1007/BF00243503.

Bigard AX, Lavier P, Ullmann L, Legrand H, Douce P, Guezennec CY: Branched-chain amino acid supplementation during repeated prolonged skiing exercises at altitude. Int J Sport Nutr. 1996, 6 (3): 295-306.

Candeloro N, Bertini I, Melchiorri G, De Lorenzo A: [Effects of prolonged administration of branched-chain amino acids on body composition and physical fitness]. Minerva Endocrinol. 1995, 20 (4): 217-23.

Stoppani J, Scheett TP, Pena J, Rudolph C, Charlebois D: Consuming a supplement containing branched-chain amino acids during a resistance-training program increases lean mass, muscle strength and fat loss. Journal of The International Society of Sport Nutrition. 2009, 6 (Suppl 1):

Wernerman J, Hammarqvist F, Vinnars E: Alpha-ketoglutarate and postoperative muscle catabolism. Lancet. 1990, 335 (8691): 701-3. 10.1016/0140-6736(90)90811-I.

Hammarqvist F, Wernerman J, Ali R, Vinnars E: Effects of an amino acid solution enriched with either branched chain amino acids or ornithine-alpha-ketoglutarate on the postoperative intracellular amino acid concentration of skeletal muscle. Br J Surg. 1990, 77 (2): 214-8. 10.1002/bjs.1800770227.

Antonio J, Stout JR: Sport Supplements. 2001, Philadelphia, PA: Lippincott, Williams and Wilkins

Mitch WE, Walser M, Sapir DG: Nitrogen sparing induced by leucine compared with that induced by its keto analogue, alpha-ketoisocaproate, in fasting obese man. J Clin Invest. 1981, 67 (2): 553-62. 10.1172/JCI110066.

Van Koevering M, Nissen S: Oxidation of leucine and alpha-ketoisocaproate to beta-hydroxy-beta-methylbutyrate in vivo. Am J Physiol. 1992, 262 (1 Pt 1): E27-31.

Slama K, Koudela K, Tenora J, Mathova A: Insect hormones in vertebrates: anabolic effects of 20-hydroxyecdysone in Japanese quail. Experientia. 1996, 52 (7): 702-6. 10.1007/BF01925578.

Slama K, Kodkoua M: Insect hormones and bioanalogues: their effect on respiratory metabolism in Dermestes vulpinus L. (Coleoptera). Biol Bull. 1975, 148 (2): 320-32. 10.2307/1540550.

Tashmukhamedova MA, Almatov KT, Syrov VN, Sultanov MB, Abidov AA: [Effect of phytoecdisteroids and anabolic steroids on liver mitochondrial respiration and oxidative phosphorylation in alloxan diabetic rats]. Nauchnye Doki Vyss Shkoly Biol Nauki. 1985 (9): 37-9.

Syrov VN: [Mechanism of the anabolic action of phytoecdisteroids in mammals]. Nauchnye Doki Vyss Shkoly Biol Nauki. 1984 (11): 16-20.

Kholodova Y: Phytoecdysteroids: biological effects, application in agriculture and complementary medicine (as presented at the 14-th Ecdysone Workshop, July, 2000, Rapperswil, Switzerland). Ukr Biokhim Zh. 2001, 73 (3): 21-9.

Toth N, Szabo A, Kacsala P, Heger J, Zador E: 20-Hydroxyecdysone increases fiber size in a muscle-specific fashion in rat. Phytomedicine. 2008, 15 (9): 691-8. 10.1016/j.phymed.2008.04.015.

Wilborn C, Taylor L, Campbell B, Kerksick C, Rasmussen C, Greenwood M, Kreider R: Effects of methoxyisoflavone, ecdysterone, and sulfo-polysaccharide supplementation on training adaptations in resistance-trained males. Journal of the International Society of Sports Nutrition. 2006, 3 (2): 10.1186/1550-2783-3-2-19.

Bowers CY: Growth hormone-releasing peptide (GHRP). Cell Mol Life Sci. 1998, 54 (12): 1316-29. 10.1007/s000180050257.

Camanni F, Ghigo E, Arvat E: Growth hormone-releasing peptides and their analogs. Front Neuroendocrinol. 1998, 19 (1): 47-72. 10.1006/frne.1997.0158.

Zachwieja JJ, Yarasheski KE: Does growth hormone therapy in conjunction with resistance exercise increase muscle force production and muscle mass in men and women aged 60 years or older? Growth hormone-releasing peptides and their analogs. Phys Ther. 1999, 79 (1): 76-82.

Coudray-Lucas C, Le Bever H, Cynober L, De Bandt JP, Carsin H: Ornithine alpha-ketoglutarate improves wound healing in severe burn patients: a prospective randomized double-blind trial versus isonitrogenous controls. Crit Care Med. 2000, 28 (6): 1772-6. 10.1097/00003246-200006000-00012.

Donati L, Ziegler F, Pongelli G, Signorini MS: Nutritional and clinical efficacy of ornithine alpha-ketoglutarate in severe burn patients. Clin Nutr. 1999, 18 (5): 307-11. 10.1016/S0261-5614(98)80029-0.

Chetlin RD, Yeater RA, Ullrich IH, Hornsby WG, Malanga CJ, Byrner RW: The effect of ornithine alpha-ketoglutarate (OKG) on healthy, weight trained men. J Exerc Physiol Online. 2000, 3 (4): [ http://faculty.css.edu/tboone2/asep/ChetlinV2.PDF ]

Brilla L, Conte V: Effects of a novel zinc-magnesium formulation on hormones and strength. J Exerc Physiol Online. 2000, 3: 26-36.

Wilborn CD, Kerksick CM, Campbell BI, Taylor LW, Marcello BM, Rasmussen CJ, Greenwood MC, Almada A, Kreider RB: Effects of Zinc Magnesium Aspartate (ZMA) Supplementation on Training Adaptations and Markers of Anabolism and Catabolism. J Int Soc Sports Nutr. 2004, 1 (2): 12-20. 10.1186/1550-2783-1-2-12.

Om AS, Chung KW: Dietary zinc deficiency alters 5 alpha-reduction and aromatization of testosterone and androgen and estrogen receptors in rat liver. J Nutr. 1996, 126 (4): 842-8.

Low SY, Taylor PM, Rennie MJ: Responses of glutamine transport in cultured rat skeletal muscle to osmotically induced changes in cell volume. J Physiol. 1996, 492: 877-85.

Rennie MJ, Ahmed A, Khogali SE, Low SY, Hundal HS, Taylor PM: Glutamine metabolism and transport in skeletal muscle and heart and their clinical relevance. J Nutr. 1996, 126 (3): 1142S-9S.

Varnier M, Leese GP, Thompson J, Rennie MJ: Stimulatory effect of glutamine on glycogen accumulation in human skeletal muscle. Am J Physiol. 1995, 269: E309-15.

Colker CM: Effects of supplemental protein on body composition and muscular strength in healthy athletic male adults. Curr Ther Res. 2000, 61 (1): 19-28. 10.1016/S0011-393X(00)88492-1.

Candow DG, Chilibeck PD, Burke DG, Davison KS, Smith-Palmer T: Effect of glutamine supplementation combined with resistance training in young adults. Eur J Appl Physiol. 2001, 86 (2): 142-9. 10.1007/s00421-001-0523-y.

Messina M: Soyfoods and soybean phyto-oestrogens (isoflavones) as possible alternatives to hormone replacement therapy (HRT). Eur J Cancer. 2000, 36 (Suppl 4): S71-2. 10.1016/S0959-8049(00)00233-1.

Messina M, Messina V: Soyfoods, soybean isoflavones, and bone health: a brief overview. J Ren Nutr. 2000, 10 (2): 63-8. 10.1016/S1051-2276(00)90001-3.

de Aloysio D, Gambacciani M, Altieri P, Ciaponi M, Ventura V, Mura M, Genazzani AR, Bottiglioni F: Bone density changes in postmenopausal women with the administration of ipriflavone alone or in association with low-dose ERT. Gynecol Endocrinol. 1997, 11 (4): 289-93. 10.3109/09513599709152548.

Slogoff S, Keats AS, Cooley DA, Reul GJ, Frazier OH, Ott DA, Duncan JM, Livesay JJ: Addition of papaverine to cardioplegia does not reduce myocardial necrosis. Ann Thorac Surg. 1986, 42 (1): 60-4.

Smart NA, McKenzie SG, Nix LM, Baldwin SE, Page K, Wade D, Hampson PK: Creatine supplementation does not improve repeat sprint performance in soccer players. Medicine & Science in Sports & Exercise. 1998, 30 (5): S140-10.1097/00005768-199805001-00794.

Aubertin-Leheudre M, Lord C, Khalil A, Dionne IJ: Six months of isoflavone supplement increases fat-free mass in obese-sarcopenic postmenopausal women: a randomized double-blind controlled trial. Eur J Clin Nutr. 2007, 61 (12): 1442-4. 10.1038/sj.ejcn.1602695.

Gonzalez-Cadavid NF, Taylor WE, Yarasheski K, Sinha-Hikim I, Ma K, Ezzat S, Shen R, Lalani R, Asa S, Mamita M, Nair G, Arver S, Bhasin S: Organization of the human myostatin gene and expression in healthy men and HIV-infected men with muscle wasting. Proc Natl Acad Sci USA. 1998, 95 (25): 14938-43. 10.1073/pnas.95.25.14938.

McPherron AC, Lawler AM, Lee SJ: Regulation of skeletal muscle mass in mice by a new TGF-beta superfamily member. Nature. 1997, 387 (6628): 83-90. 10.1038/387083a0.

McPherron AC, Lee SJ: Double muscling in cattle due to mutations in the myostatin gene. Proc Natl Acad Sci USA. 1997, 94 (23): 12457-61. 10.1073/pnas.94.23.12457.

Grobet L, Martin LJ, Poncelet D, Pirottin D, Brouwers B, Riquet J, Schoeberlein A, Dunner S, Menissier F, Massabanda J, Fries R, Hanset R, Georges M: A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle. Nat Genet. 1997, 17 (1): 71-4. 10.1038/ng0997-71.

Kambadur R, Sharma M, Smith TP, Bass JJ: Mutations in myostatin (GDF8) in double-muscled Belgian Blue and Piedmontese cattle. Genome Res. 1997, 7 (9): 910-6.

Ivey FM, Roth SM, Ferrell RE, Tracy BL, Lemmer JT, Hurlbut DE, Martel GF, Siegel EL, Fozard JL, Jeffrey Metter E, Fleg JL, Hurley BF: Effects of age, gender, and myostatin genotype on the hypertrophic response to heavy resistance strength training. J Gerontol A Biol Sci Med Sci. 2000, 55 (11): M641-8.

Carlson CJ, Booth FW, Gordon SE: Skeletal muscle myostatin mRNA expression is fiber-type specific and increases during hindlimb unloading. Am J Physiol. 1999, 277 (2 Pt 2): R601-6.

Willoughby DS: Effects of an alleged myostatin-binding supplement and heavy resistance training on serum myostatin, muscle strength and mass, and body composition. Int J Sport Nutr Exerc Metab. 2004, 14 (4): 461-72.

Saremi A, Gharakhanloo R, Sharghi S, Gharaati MR, Larijani B, Omidfar K: Effects of oral creatine and resistance training on serum myostatin and GASP-1. Mol Cell Endocrinol. 2009

Green NR, Ferrando AA: Plasma boron and the effects of boron supplementation in males. Environ Health Perspect. 1994, 102 (Suppl 7): 73-7. 10.2307/3431966.

Ferrando AA, Green NR: The effect of boron supplementation on lean body mass, plasma testosterone levels, and strength in male bodybuilders. Int J Sport Nutr. 1993, 3 (2): 140-9.

Evans GW: The effect of chromium picolinate on insulin controlled parameters in humans. Int Biosc Med Res. 1989, 11: 163-80.

Hasten DL, Rome EP, Franks BD, Hegsted M: Effects of chromium picolinate on beginning weight training students. Int J Sport Nutr. 1992, 2 (4): 343-50.

Grant KE, Chandler RM, Castle AL, Ivy JL: Chromium and exercise training: effect on obese women. Med Sci Sports Exerc. 1997, 29 (8): 992-8.

Campbell WW, Joseph LJ, Anderson RA, Davey SL, Hinton J, Evans WJ: Effects of resistive training and chromium picolinate on body composition and skeletal muscle size in older women. Int J Sport Nutr Exerc Metab. 2002, 12 (2): 125-35.

Campbell WW, Joseph LJ, Davey SL, Cyr-Campbell D, Anderson RA, Evans WJ: Effects of resistance training and chromium picolinate on body composition and skeletal muscle in older men. J Appl Physiol. 1999, 86 (1): 29-39.

Walker LS, Bemben MG, Bemben DA, Knehans AW: Chromium picolinate effects on body composition and muscular performance in wrestlers. Med Sci Sports Exerc. 1998, 30 (12): 1730-7. 10.1097/00005768-199812000-00012.

Livolsi JM, Adams GM, Laguna PL: The effect of chromium picolinate on muscular strength and body composition in women athletes. J Strength Cond Res. 2001, 15 (2): 161-6. 10.1519/1533-4287(2001)015<0161:TEOCPO>2.0.CO;2.

Volpe SL, Huang HW, Larpadisorn K, Lesser II: Effect of chromium supplementation and exercise on body composition, resting metabolic rate and selected biochemical parameters in moderately obese women following an exercise program. J Am Coll Nutr. 2001, 20 (4): 293-306.

Hallmark MA, Reynolds TH, DeSouza CA, Dotson CO, Anderson RA, Rogers MA: Effects of chromium and resistive training on muscle strength and body composition. Med Sci Sports Exerc. 1996, 28 (1): 139-44. 10.1097/00005768-199601000-00025.

Lukaski HC, Bolonchuk WW, Siders WA, Milne DB: Chromium supplementation and resistance training: effects on body composition, strength, and trace element status of men. Am J Clin Nutr. 1996, 63 (6): 954-65.

Clancy SP, Clarkson PM, DeCheke ME, Nosaka K, Freedson PS, Cunningham JJ, Valentine B: Effects of chromium picolinate supplementation on body composition, strength, and urinary chromium loss in football players. Int J Sport Nutr. 1994, 4 (2): 142-53.

Pariza MW, Park Y, Cook ME: Conjugated linoleic acid and the control of cancer and obesity. Toxicol Sci. 1999, 52 (2 Suppl): 107-l10.

Pariza MW, Park Y, Cook ME: Mechanisms of action of conjugated linoleic acid: evidence and speculation. Proc Soc Exp Biol Med. 2000, 223 (1): 8-13. 10.1046/j.1525-1373.2000.22302.x.

Pariza MW, Park Y, Cook ME: The biologically active isomers of conjugated linoleic acid. Prog Lipid Res. 2001, 40 (4): 283-98. 10.1016/S0163-7827(01)00008-X.

DeLany JP, Blohm F, Truett AA, Scimeca JA, West DB: Conjugated linoleic acid rapidly reduces body fat content in mice without affecting energy intake. Am J Physiol. 1999, 276 (4 Pt 2): R1172-9.

DeLany JP, West DB: Changes in body composition with conjugated linoleic acid. J Am Coll Nutr. 2000, 19 (4): 487S-93S.

Park Y, Albright KJ, Liu W, Storkson JM, Cook ME, Pariza MW: Effect of conjugated linoleic acid on body composition in mice. Lipids. 1997, 32 (8): 853-8. 10.1007/s11745-997-0109-x.

Blankson H, Stakkestad JA, Fagertun H, Thom E, Wadstein J, Gudmundsen O: Conjugated linoleic acid reduces body fat mass in overweight and obese humans. J Nutr. 2000, 130 (12): 2943-8.

Gaullier JM, Berven G, Blankson H, Gudmundsen O: Clinical trial results support a preference for using CLA preparations enriched with two isomers rather than four isomers in human studies. Lipids. 2002, 37 (11): 1019-25. 10.1007/s11745-002-0995-y.

Pinkoski C, Chilibeck PD, Candow DG, Esliger D, Ewaschuk JB, Facci M, Farthing JP, Zello GA: The effects of conjugated linoleic acid supplementation during resistance training. Med Sci Sports Exerc. 2006, 38 (2): 339-48. 10.1249/01.mss.0000183860.42853.15.

Tarnopolsky M, Zimmer A, Paikin J, Safdar A, Aboud A, Pearce E, Roy B, Doherty T: Creatine monohydrate and conjugated linoleic acid improve strength and body composition following resistance exercise in older adults. PLoS One. 2007, 2 (10): e991-10.1371/journal.pone.0000991.

Campbell B, Kreider RB: Conjugated linoleic acids. Curr Sports Med Rep. 2008, 7 (4): 237-41.

Wheeler KB, Garleb KA: Gamma oryzanol-plant sterol supplementation: metabolic, endocrine, and physiologic effects. Int J Sport Nutr. 1991, 1 (2): 170-7.

Fry AC, Bonner E, Lewis DL, Johnson RL, Stone MH, Kraemer WJ: The effects of gamma-oryzanol supplementation during resistance exercise training. Int J Sport Nutr. 1997, 7 (4): 318-29.

Bhasin S, Woodhouse L, Casaburi R, Singh AB, Mac RP, Lee M, Yarasheski KE, Sinha-Hikim I, Dzekov C, Dzekov J, Magliano L, Storer TW: Older men are as responsive as young men to the anabolic effects of graded doses of testosterone on the skeletal muscle. J Clin Endocrinol Metab. 2005, 90 (2): 678-88. 10.1210/jc.2004-1184.

Kuhn CM: Anabolic steroids. Recent Prog Horm Res. 2002, 57: 411-34. 10.1210/rp.57.1.411.

Limbird TJ: Anabolic steroids in the training and treatment of athletes. Compr Ther. 1985, 11 (1): 25-30.

Lukas SE: Current perspectives on anabolic-androgenic steroid abuse. Trends Pharmacol Sci. 1993, 14 (2): 61-8. 10.1016/0165-6147(93)90032-F.

Sattler FR, Castaneda-Sceppa C, Binder EF, Schroeder ET, Wang Y, Bhasin S, Kawakubo M, Stewart Y, Yarasheski KE, Ulloor J, Colletti P, Roubenoff R, Azen SP: Testosterone and growth hormone improve body composition and muscle performance in older men. J Clin Endocrinol Metab. 2009, 94 (6): 1991-2001. 10.1210/jc.2008-2338.

Storer TW, Woodhouse L, Magliano L, Singh AB, Dzekov C, Dzekov J, Bhasin S: Changes in muscle mass, muscle strength, and power but not physical function are related to testosterone dose in healthy older men. J Am Geriatr Soc. 2008, 56 (11): 1991-9. 10.1111/j.1532-5415.2008.01927.x.

Wagner JC: Enhancement of athletic performance with drugs. An overview. Sports Med. 1991, 12 (4): 250-65. 10.2165/00007256-199112040-00004.

Yarasheski KE: Growth hormone effects on metabolism, body composition, muscle mass, and strength. Exerc Sport Sci Rev. 1994, 22: 285-312. 10.1249/00003677-199401000-00013.

Smart T: Other therapies for wasting. GMHC Treat Issues. 1995, 9 (5): 7-8. 12

Casaburi R: Skeletal muscle dysfunction in chronic obstructive pulmonary disease. Med Sci Sports Exerc. 2001, 33 (7 Suppl): S662-70.

Hayes VY, Urban RJ, Jiang J, Marcell TJ, Helgeson K, Mauras N: Recombinant human growth hormone and recombinant human insulin-like growth factor I diminish the catabolic effects of hypogonadism in man: metabolic and molecular effects. J Clin Endocrinol Metab. 2001, 86 (5): 2211-9. 10.1210/jc.86.5.2211.

Newshan G, Leon W: The use of anabolic agents in HIV disease. Int J STD AIDS. 2001, 12 (3): 141-4. 10.1258/0956462011916893.

Tenover JS: Androgen replacement therapy to reverse and/or prevent age-associated sarcopenia in men. Baillieres Clin Endocrinol Metab. 1998, 12 (3): 419-25. 10.1016/S0950-351X(98)80153-5.

Bross R, Casaburi R, Storer TW, Bhasin S: Androgen effects on body composition and muscle function: implications for the use of androgens as anabolic agents in sarcopenic states. Baillieres Clin Endocrinol Metab. 1998, 12 (3): 365-78. 10.1016/S0950-351X(98)80077-3.

Casaburi R: Rationale for anabolic therapy to facilitate rehabilitation in chronic obstructive pulmonary disease. Baillieres Clin Endocrinol Metab. 1998, 12 (3): 407-18. 10.1016/S0950-351X(98)80134-1.

Johansen KL, Mulligan K, Schambelan M: Anabolic effects of nandrolone decanoate in patients receiving dialysis: a randomized controlled trial. Jama. 1999, 281 (14): 1275-81. 10.1001/jama.281.14.1275.

Sattler FR, Jaque SV, Schroeder ET, Olson C, Dube MP, Martinez C, Briggs W, Horton R, Azen S: Effects of pharmacological doses of nandrolone decanoate and progressive resistance training in immunodeficient patients infected with human immunodeficiency virus. J Clin Endocrinol Metab. 1999, 84 (4): 1268-76. 10.1210/jc.84.4.1268.

Beiner JM, Jokl P, Cholewicki J, Panjabi MM: The effect of anabolic steroids and corticosteroids on healing of muscle contusion injury. Am J Sports Med. 1999, 27 (1): 2-9.

Ferreira IM, Verreschi IT, Nery LE, Goldstein RS, Zamel N, Brooks D, Jardim JR: The influence of 6 months of oral anabolic steroids on body mass and respiratory muscles in undernourished COPD patients. Chest. 1998, 114 (1): 19-28. 10.1378/chest.114.1.19.

Bhasin S, Bremner WJ: Clinical review 85: Emerging issues in androgen replacement therapy. J Clin Endocrinol Metab. 1997, 82 (1): 3-8. 10.1210/jc.82.1.3.

Hoffman JR, Kraemer WJ, Bhasin S, Storer T, Ratamess NA, Haff GG, Willoughby DS, Rogol AD: Position stand on androgen and human growth hormone use. J Strength Cond Res. 2009, 23 (5 Suppl): S1-S59.

Ferrando AA, Sheffield-Moore M, Paddon-Jones D, Wolfe RR, Urban RJ: Differential anabolic effects of testosterone and amino acid feeding in older men. J Clin Endocrinol Metab. 2003, 88 (1): 358-62. 10.1210/jc.2002-021041.

Meeuwsen IB, Samson MM, Duursma SA, Verhaar HJ: Muscle strength and tibolone: a randomised, double-blind, placebo-controlled trial. Bjog. 2002, 109 (1): 77-84.

King DS, Sharp RL, Vukovich MD, Brown GA, Reifenrath TA, Uhl NL, Parsons KA: Effect of oral androstenedione on serum testosterone and adaptations to resistance training in young men: a randomized controlled trial. Jama. 1999, 281 (21): 2020-8. 10.1001/jama.281.21.2020.

Carter WJ: Effect of anabolic hormones and insulin-like growth factor-I on muscle mass and strength in elderly persons. Clin Geriatr Med. 1995, 11 (4): 735-48.

Soe M, Jensen KL, Gluud C: [The effect of anabolic androgenic steroids on muscle strength, body weight and lean body mass in body-building men]. Ugeskr Laeger. 1989, 151 (10): 610-3.

Griggs RC, Pandya S, Florence JM, Brooke MH, Kingston W, Miller JP, Chutkow J, Herr BE, Moxley RT: Randomized controlled trial of testosterone in myotonic dystrophy. Neurology. 1989, 39 (2 Pt 1): 219-22.

Crist DM, Stackpole PJ, Peake GT: Effects of androgenic-anabolic steroids on neuromuscular power and body composition. J Appl Physiol. 1983, 54 (2): 366-70.

Ward P: The effect of an anabolic steroid on strength and lean body mass. Med Sci Sports. 1973, 5 (4): 277-82.

Varriale P, Mirzai-tehrane M, Sedighi A: Acute myocardial infarction associated with anabolic steroids in a young HIV-infected patient. Pharmacotherapy. 1999, 19 (7): 881-4. 10.1592/phco.19.10.881.31552.

Kibble MW, Ross MB: Adverse effects of anabolic steroids in athletes. Clin Pharm. 1987, 6 (9): 686-92.

Gruber AJ, Pope HG: Psychiatric and medical effects of anabolic-androgenic steroid use in women. Psychother Psychosom. 2000, 69 (1): 19-26. 10.1159/000012362.

Lamb DR: Anabolic steroids in athletics: how well do they work and how dangerous are they? Am. J Sports Med. 1984, 12 (1): 31-8. 10.1177/036354658401200105.

Salke RC, Rowland TW, Burke EJ: Left ventricular size and function in body builders using anabolic steroids. Med Sci Sports Exerc. 1985, 17 (6): 701-4. 10.1249/00005768-198512000-00014.

Brown GA, Martini ER, Roberts BS, Vukovich MD, King DS: Acute hormonal response to sublingual androstenediol intake in young men. J Appl Physiol. 2002, 92 (1): 142-6.

Brown GA, McKenzie D: Acute resistance exercise does not change the hormonal response to sublingual androstenediol intake. Eur J Appl Physiol. 2006, 97 (4): 404-12. 10.1007/s00421-006-0194-9.

Broeder CE, Quindry J, Brittingham K, Panton L, Thomson J, Appakondu S, Breuel K, Byrd R, Douglas J, Earnest C, Mitchell C, Olson M, Roy T, Yarlagadda C: The Andro Project: physiological and hormonal influences of androstenedione supplementation in men 35 to 65 years old participating in a high-intensity resistance training program. Arch Intern Med. 2000, 160 (20): 3093-104. 10.1001/archinte.160.20.3093.

Ballantyne CS, Phillips SM, MacDonald JR, Tarnopolsky MA, MacDougall JD: The acute effects of androstenedione supplementation in healthy young males. Can J Appl Physiol. 2000, 25 (1): 68-78.

Brown GA, Vukovich MD, Sharp RL, Reifenrath TA, Parsons KA, King DS: Effect of oral DHEA on serum testosterone and adaptations to resistance training in young men. J Appl Physiol. 1999, 87 (6): 2274-83.

van Gammeren D, Falk D, Antonio J: Effects of norandrostenedione and norandrostenediol in resistance-trained men. Nutrition. 2002, 18 (9): 734-7. 10.1016/S0899-9007(02)00834-1.

Van Gammeren D, Falk D, Antonio J: The effects of supplementation with 19-nor-4-androstene-3,17-dione and 19-nor-4-androstene-3,17-diol on body composition and athletic performance in previously weight-trained male athletes. Eur J Appl Physiol. 2001, 84 (5): 426-31. 10.1007/s004210100395.

Pipe A: Effects of testosterone precursor supplementation on intensive weight training. Clin J Sport Med. 2001, 11 (2): 126-10.1097/00042752-200104000-00014.

Mauras N, Lima J, Patel D, Rini A, di Salle E, Kwok A, Lippe B: Pharmacokinetics and dose finding of a potent aromatase inhibitor, aromasin (exemestane), in young males. J Clin Endocrinol Metab. 2003, 88 (12): 5951-6. 10.1210/jc.2003-031279.

Rohle D, Wilborn C, Taylor L, Mulligan C, Kreider R, Willoughby D: Effects of eight weeks of an alleged aromatase inhibiting nutritional supplement 6-OXO (androst-4-ene-3,6,17-trione) on serum hormone profiles and clinical safety markers in resistance-trained, eugonadal males. J Int Soc Sports Nutr. 2007, 4: 13-10.1186/1550-2783-4-13.

Willoughby DS, Wilborn C, Taylor L, Campbell W: Eight weeks of aromatase inhibition using the nutritional supplement Novedex XT: effects in young, eugonadal men. Int J Sport Nutr Exerc Metab. 2007, 17 (1): 92-108.

Antonio J, Uelmen J, Rodriguez R, Earnest C: The effects of Tribulus terrestris on body composition and exercise performance in resistance-trained males. Int J Sport Nutr Exerc Metab. 2000, 10 (2): 208-15.

Brown GA, Vukovich MD, Martini ER, Kohut ML, Franke WD, Jackson DA, King DS: Effects of androstenedione-herbal supplementation on serum sex hormone concentrations in 30- to 59-year-old men. Int J Vitam Nutr Res. 2001, 71 (5): 293-301. 10.1024/0300-9831.71.5.293.

Rogerson S, Riches CJ, Jennings C, Weatherby RP, Meir RA, Marshall-Gradisnik SM: The effect of five weeks of Tribulus terrestris supplementation on muscle strength and body composition during preseason training in elite rugby league players. J Strength Cond Res. 2007, 21 (2): 348-53. 10.1519/R-18395.1.

Cohen N, Halberstam M, Shlimovich P, Chang CJ, Shamoon H, Rossetti L: Oral vanadyl sulfate improves hepatic and peripheral insulin sensitivity in patients with non-insulin-dependent diabetes mellitus. J Clin Invest. 1995, 95 (6): 2501-9. 10.1172/JCI117951.

Boden G, Chen X, Ruiz J, van Rossum GD, Turco S: Effects of vanadyl sulfate on carbohydrate and lipid metabolism in patients with non-insulin-dependent diabetes mellitus. Metabolism. 1996, 45 (9): 1130-5. 10.1016/S0026-0495(96)90013-X.

Halberstam M, Cohen N, Shlimovich P, Rossetti L, Shamoon H: Oral vanadyl sulfate improves insulin sensitivity in NIDDM but not in obese nondiabetic subjects. Diabetes. 1996, 45 (5): 659-66. 10.2337/diabetes.45.5.659.

Fawcett JP, Farquhar SJ, Walker RJ, Thou T, Lowe G, Goulding A: The effect of oral vanadyl sulfate on body composition and performance in weight-training athletes. Int J Sport Nutr. 1996, 6 (4): 382-90.

Fawcett JP, Farquhar SJ, Thou T, Shand BI: Oral vanadyl sulphate does not affect blood cells, viscosity or biochemistry in humans. Pharmacol Toxicol. 1997, 80 (4): 202-6. 10.1111/j.1600-0773.1997.tb00397.x.

Kreider R: New weight-control options. Func Foods Nutraceut. 2002, 34-42.

Hoie LH, Bruusgaard D, Thom E: Reduction of body mass and change in body composition on a very low calorie diet. Int J Obes Relat Metab Disord. 1993, 17 (1): 17-20.

Bryner RW, Ullrich IH, Sauers J, Donley D, Hornsby G, Kolar M, Yeater R: Effects of resistance vs. aerobic training combined with an 800 calorie liquid diet on lean body mass and resting metabolic rate. J Am Coll Nutr. 1999, 18 (2): 115-21.

Meckling KA, Sherfey R: A randomized trial of a hypocaloric high-protein diet, with and without exercise, on weight loss, fitness, and markers of the Metabolic Syndrome in overweight and obese women. Appl Physiol Nutr Metab. 2007, 32 (4): 743-52. 10.1139/H07-059.

Aoyama T, Fukui K, Takamatsu K, Hashimoto Y, Yamamoto T: Soy protein isolate and its hydrolysate reduce body fat of dietary obese rats and genetically obese mice (yellow KK). Nutrition. 2000, 16 (5): 349-54. 10.1016/S0899-9007(00)00230-6.

Baba NH, Sawaya S, Torbay N, Habbal Z, Azar S, Hashim SA: High protein vs high carbohydrate hypoenergetic diet for the treatment of obese hyperinsulinemic subjects. Int J Obes Relat Metab Disord. 1999, 23 (11): 1202-6. 10.1038/sj.ijo.0801064.

Clifton P: High protein diets and weight control. Nutr Metab Cardiovasc Dis. 2009, 19 (6): 379-82. 10.1016/j.numecd.2009.02.011.

Heymsfield SB, van Mierlo CA, Knaap van der HC, Heo M, Frier HI: Weight management using a meal replacement strategy: meta and pooling analysis from six studies. Int J Obes Relat Metab Disord. 2003, 27 (5): 537-49. 10.1038/sj.ijo.0802258.

Skov AR, Toubro S, Ronn B, Holm L, Astrup A: Randomized trial on protein vs carbohydrate in ad libitum fat reduced diet for the treatment of obesity. Int J Obes Relat Metab Disord. 1999, 23 (5): 528-36. 10.1038/sj.ijo.0800867.

Toubro S, Astrup AV: [A randomized comparison of two weight-reducing diets. Calorie counting versus low-fat carbohydrate-rich ad libitum diet]. Ugeskr Laeger. 1998, 160 (6): 816-20.

Wal JS, McBurney MI, Cho S, Dhurandhar NV: Ready-to-eat cereal products as meal replacements for weight loss. Int J Food Sci Nutr. 2007, 58 (5): 331-40. 10.1080/09637480701240802.

Reaven GM: Diet and Syndrome X. Curr Atheroscler Rep. 2000, 2 (6): 503-7. 10.1007/s11883-000-0050-z.

Treyzon L, Chen S, Hong K, Yan E, Carpenter CL, Thames G, Bowerman S, Wang HJ, Elashoff R, Li Z: A controlled trial of protein enrichment of meal replacements for weight reduction with retention of lean body mass. Nutr J. 2008, 7: 23-10.1186/1475-2891-7-23.

Hasani-Ranjbar S, Nayebi N, Larijani B, Abdollahi M: A systematic review of the efficacy and safety of herbal medicines used in the treatment of obesity. World J Gastroenterol. 2009, 15 (25): 3073-85. 10.3748/wjg.15.3073.

Greenway FL, De Jonge L, Blanchard D, Frisard M, Smith SR: Effect of a dietary herbal supplement containing caffeine and ephedra on weight, metabolic rate, and body composition. Obes Res. 2004, 12 (7): 1152-7. 10.1038/oby.2004.144.

Coffey CS, Steiner D, Baker BA, Allison DB: A randomized double-blind placebo-controlled clinical trial of a product containing ephedrine, caffeine, and other ingredients from herbal sources for treatment of overweight and obesity in the absence of lifestyle treatment. Int J Obes Relat Metab Disord. 2004, 28 (11): 1411-9. 10.1038/sj.ijo.0802784.

Boozer CN, Daly PA, Homel P, Solomon JL, Blanchard D, Nasser JA, Strauss R, Meredith T: Herbal ephedra/caffeine for weight loss: a 6-month randomized safety and efficacy trial. Int J Obes Relat Metab Disord. 2002, 26 (5): 593-604. 10.1038/sj.ijo.0802023.

Boozer C, Nasser J, SB H, Wang V, Chen G, Solomon J: An herbal supplement containing Ma Huang-Guarana for weight loss: a randomized, double-blind trial. Int J Obes Relat Metab Disord. 2001, 25: 316-24. 10.1038/sj.ijo.0801539.

Boozer C, Daly P, Homel P, Solomon J, Blanchard D, Nasser J, Strauss R, Merideth T: Herbal ephedra/caffeine for weight loss: a 6-month randomized safety and efficacy trial. Int J Obesity. 2002, 26: 593-604. 10.1038/sj.ijo.0802023.

Molnar D, Torok K, Erhardt E, Jeges S: Safety and efficacy of treatment with an ephedrine/caffeine mixture. The first double-blind placebo-controlled pilot study in adolescents. Int J Obes Relat Metab Disord. 2000, 24 (12): 1573-8. 10.1038/sj.ijo.0801433.

Molnar D: Effects of ephedrine and aminophylline on resting energy expenditure in obese adolescents. Int J Obes Relat Metab Disord. 1993, 17 (Suppl 1): S49-52.

Greenway FL: The safety and efficacy of pharmaceutical and herbal caffeine and ephedrine use as a weight loss agent. Obes Rev. 2001, 2 (3): 199-211. 10.1046/j.1467-789x.2001.00038.x.

Greenway F, Raum W, DeLany J: The effect of an herbal dietary supplement containing ephedrine and caffeine on oxygen consumption in humans. J Altern Complement Med. 2000, 6 (6): 553-5. 10.1089/acm.2000.6.553.

Greenway F, Herber D, Raum W, Morales S: Double-blind, randomized, placebo-controlled clinical trials with non-prescription medications for the treatment of obesity. Obes Res. 1999, 7 (4): 370-8.

Greenway FL, Ryan DH, Bray GA, Rood JC, Tucker EW, Smith SR: Pharmaceutical cost savings of treating obesity with weight loss medications. Obes Res. 1999, 7 (6): 523-31.

Hackman RM, Havel PJ, Schwartz HJ, Rutledge JC, Watnik MR, Noceti EM, Stohs SJ, Stern JS, Keen CL: Multinutrient supplement containing ephedra and caffeine causes weight loss and improves metabolic risk factors in obese women: a randomized controlled trial. Int J Obes (Lond). 2006, 30 (10): 1545-56. 10.1038/sj.ijo.0803283.

Bent S, Tiedt T, Odden M, Shlipak M: The relative safety of ephedra compared with other herbal products. Ann Intern Med. 2003, 138: 468-471.

Fleming GA: The FDA, regulation, and the risk of stroke. N Engl J Med. 2000, 343 (25): 1886-7. 10.1056/NEJM200012213432510.

Anderson JW, Baird P, Davis RH, Ferreri S, Knudtson M, Koraym A, Waters V, Williams CL: Health benefits of dietary fiber. Nutr Rev. 2009, 67 (4): 188-205. 10.1111/j.1753-4887.2009.00189.x.

Shai I, Schwarzfuchs D, Henkin Y, Shahar DR, Witkow S, Greenberg I, Golan R, Fraser D, Bolotin A, Vardi H, Tangi-Rozental O, Zuk-Ramot R, Sarusi B, Brickner D, Schwartz Z, Sheiner E, Marko R, Katorza E, Thiery J, Fiedler GM, Bluher M, Stumvoll M, Stampfer MJ: Weight loss with a low-carbohydrate, Mediterranean, or low-fat diet. N Engl J Med. 2008, 359 (3): 229-41. 10.1056/NEJMoa0708681.

Raben A, Jensen ND, Marckmann P, Sandstrom B, Astrup AV: [Spontaneous weight loss in young subjects of normal weight after 11 weeks of unrestricted intake of a low-fat/high-fiber diet]. Ugeskr Laeger. 1997, 159 (10): 1448-53.

Melanson KJ, Angelopoulos TJ, Nguyen VT, Martini M, Zukley L, Lowndes J, Dube TJ, Fiutem JJ, Yount BW, Rippe JM: Consumption of whole-grain cereals during weight loss: effects on dietary quality, dietary fiber, magnesium, vitamin B-6, and obesity. J Am Diet Assoc. 2006, 106 (9): 1380-8. 10.1016/j.jada.2006.06.003. quiz 9-90

Nieman DC, Cayea EJ, Austin MD, Henson DA, McAnulty SR, Jin F: Chia seed does not promote weight loss or alter disease risk factors in overweight adults. Nutr Res. 2009, 29 (6): 414-8. 10.1016/j.nutres.2009.05.011.

Saltzman E, Moriguti JC, Das SK, Corrales A, Fuss P, Greenberg AS, Roberts SB: Effects of a cereal rich in soluble fiber on body composition and dietary compliance during consumption of a hypocaloric diet. J Am Coll Nutr. 2001, 20 (1): 50-7.

Sartorelli DS, Franco LJ, Cardoso MA: High intake of fruits and vegetables predicts weight loss in Brazilian overweight adults. Nutr Res. 2008, 28 (4): 233-8. 10.1016/j.nutres.2008.02.004.

Barr SI: Increased dairy product or calcium intake: is body weight or composition affected in humans?. J Nutr. 2003, 133 (1): 245S-8S.

Lanou AJ, Barnard ND: Dairy and weight loss hypothesis: an evaluation of the clinical trials. Nutr Rev. 2008, 66 (5): 272-9. 10.1111/j.1753-4887.2008.00032.x.

Menon VB, Baxmann AC, Froeder L, Martini LA, Heilberg IP: Effects of calcium supplementation on body weight reduction in overweight calcium stone formers. Urol Res. 2009, 37 (3): 133-9. 10.1007/s00240-009-0187-3.

Shapses SA, Heshka S, Heymsfield SB: Effect of calcium supplementation on weight and fat loss in women. J Clin Endocrinol Metab. 2004, 89 (2): 632-7. 10.1210/jc.2002-021136.

Wagner G, Kindrick S, Hertzler S, DiSilvestro RA: Effects of various forms of calcium on body weight and bone turnover markers in women participating in a weight loss program. J Am Coll Nutr. 2007, 26 (5): 456-61.

Yanovski JA, Parikh SJ, Yanoff LB, Denkinger BI, Calis KA, Reynolds JC, Sebring NG, McHugh T: Effects of calcium supplementation on body weight and adiposity in overweight and obese adults: a randomized trial. Ann Intern Med. 2009, 150 (12): 821-9. W145-6

Zemel M, Thompson W, Zemel P, Nocton A, Milstead A, Morris K, Campbell P: Dietary calcium and dairy products accelerate weight and fat-loss during energy restriction in obese adults. Clin Nutri. 2002, 75-

Zemel MB: Role of dietary calcium and dairy products in modulating adiposity. Lipids. 2003, 38 (2): 139-46. 10.1007/s11745-003-1044-6.

Zemel MB: Regulation of adiposity and obesity risk by dietary calcium: mechanisms and implications. J Am Coll Nutr. 2002, 21 (2): 146S-51S.

Zemel MB: Mechanisms of dairy modulation of adiposity. J Nutr. 2003, 133 (1): 252S-6S.

Zemel MB, Shi H, Greer B, Dirienzo D, Zemel PC: Regulation of adiposity by dietary calcium. Faseb J. 2000, 14 (9): 1132-8.

Davies KM, Heaney RP, Recker RR, Lappe JM, Barger-Lux MJ, Rafferty K, Hinders S: Calcium intake and body weight. J Clin Endocrinol Metab. 2000, 85 (12): 4635-8. 10.1210/jc.85.12.4635.

Sarma DN, Barrett ML, Chavez ML, Gardiner P, Ko R, Mahady GB, Marles RJ, Pellicore LS, Giancaspro GI, Low Dog T: Safety of green tea extracts: a systematic review by the US Pharmacopeia. Drug Saf. 2008, 31 (6): 469-84. 10.2165/00002018-200831060-00003.

Nagle DG, Ferreira D, Zhou YD: Epigallocatechin-3-gallate (EGCG): chemical and biomedical perspectives. Phytochemistry. 2006, 67 (17): 1849-55. 10.1016/j.phytochem.2006.06.020.

Shixian Q, VanCrey B, Shi J, Kakuda Y, Jiang Y: Green tea extract thermogenesis-induced weight loss by epigallocatechin gallate inhibition of catechol-O-methyltransferase. J Med Food. 2006, 9 (4): 451-8. 10.1089/jmf.2006.9.451.

Nakagawa K, Ninomiya M, Okubo T, Aoi N, Juneja LR, Kim M, Yamanaka K, Miyazawa T: Tea catechin supplementation increases antioxidant capacity and prevents phospholipid hydroperoxidation in plasma of humans. J Agric Food Chem. 1999, 47 (10): 3967-73. 10.1021/jf981195l.

Dulloo A, Duret C, Rohrer D, Girardier L, Mensi N, Fathi M, Chantre P, Vandermander J: Efficacy of a green tea extract rich in catechin polyphenols and caffeine in increasing 24-h energy expenditure and fat oxidation in humans. Am J Clin Nutr. 2000, 70 (6): 1040-5.

Dulloo AG, Duret C, Rohrer D, Girardier L, Mensi N, Fathi M, Chantre P, Vandermander J: Efficacy of a green tea extract rich in catechin polyphenols and caffeine in increasing 24-h energy expenditure and fat oxidation in humans. Am J Clin Nutr. 1999, 70 (6): 1040-5.

Di Pierro F, Menghi AB, Barreca A, Lucarelli M, Calandrelli A: Greenselect Phytosome as an adjunct to a low-calorie diet for treatment of obesity: a clinical trial. Altern Med Rev. 2009, 14 (2): 154-60.

Maki KC, Reeves MS, Farmer M, Yasunaga K, Matsuo N, Katsuragi Y, Komikado M, Tokimitsu I, Wilder D, Jones F, Blumberg JB, Cartwright Y: Green tea catechin consumption enhances exercise-induced abdominal fat loss in overweight and obese adults. J Nutr. 2009, 139 (2): 264-70.

Fallon E, Zhong L, Furne JK, Levitt M: A mixture of extracts of black and green teas and mulberry leaf did not reduce weight gain in rats fed a high-fat diet. Altern Med Rev. 2008, 13 (1): 43-9.

Hsu CH, Tsai TH, Kao YH, Hwang KC, Tseng TY, Chou P: Effect of green tea extract on obese women: a randomized, double-blind, placebo-controlled clinical trial. Clin Nutr. 2008, 27 (3): 363-70. 10.1016/j.clnu.2008.03.007.

MacDonald HB: Conjugated linoleic acid and disease prevention: a review of current knowledge. J Am Coll Nutr. 2000, 19 (2 Suppl): 111S-8S.

Park Y, Albright KJ, Storkson JM, Liu W, Cook ME, Pariza MW: Changes in body composition in mice during feeding and withdrawal of conjugated linoleic acid. Lipids. 1999, 34 (3): 243-8. 10.1007/s11745-999-0359-7.

Colakoglu S, Colakoglu M, Taneli F, Cetinoz F, Turkmen M: Cumulative effects of conjugated linoleic acid and exercise on endurance development, body composition, serum leptin and insulin levels. J Sports Med Phys Fitness. 2006, 46 (4): 570-7.

Lowery LM, Appicelli PA, PWR L: Conjugated linoleic acid enhances muscle size and strength gains in novice bodybuilders. Med Sci Sports Exerc. 1998, 30 (5): S182-

Riserus U, Arner P, Brismar K, Vessby B: Treatment with dietary trans10cis12 conjugated linoleic acid causes isomer-specific insulin resistance in obese men with the metabolic syndrome. Diabetes Care. 2002, 25 (9): 1516-21. 10.2337/diacare.25.9.1516.

Riserus U, Basu S, Jovinge S, Fredrikson GN, Arnlov J, Vessby B: Supplementation with conjugated linoleic acid causes isomer-dependent oxidative stress and elevated C-reactive protein: a potential link to fatty acid-induced insulin resistance. Circulation. 2002, 106 (15): 1925-9. 10.1161/01.CIR.0000033589.15413.48.

Riserus U, Berglund L, Vessby B: Conjugated linoleic acid (CLA) reduced abdominal adipose tissue in obese middle-aged men with signs of the metabolic syndrome: a randomised controlled trial. Int J Obes Relat Metab Disord. 2001, 25 (8): 1129-35. 10.1038/sj.ijo.0801659.

Thom E, Wadstein J, Gudmundsen O: Conjugated linoleic acid reduces body fat in healthy exercising humans. J Int Med Res. 2001, 29 (5): 392-6.

Cornish SM, Candow DG, Jantz NT, Chilibeck PD, Little JP, Forbes S, Abeysekara S, Zello GA: Conjugated linoleic acid combined with creatine monohydrate and whey protein supplementation during strength training. Int J Sport Nutr Exerc Metab. 2009, 19 (1): 79-96.

Beuker F, Haak H, Schwietz H, editors: CLA and body styling. Symposium: Vitamine und Zusatzstoffe; Jena (Thhr.). 1999

Kreider RB, Ferreira MP, Greenwood M, Wilson M, Almada AL: Effects of conjugated linoleic acid supplementation during resistance training on body composition, bone density, strength, and selected hematological markers. J Strength Cond Res. 2002, 16 (3): 325-34. 10.1519/1533-4287(2002)016<0325:EOCLAS>2.0.CO;2.

Malpuech-Brugere C, Verboeket-van de Venne WP, Mensink RP, Arnal MA, Morio B, Brandolini M, Saebo A, Lassel TS, Chardigny JM, Sebedio JL, Beaufrere B: Effects of two conjugated linoleic Acid isomers on body fat mass in overweight humans. Obes Res. 2004, 12 (4): 591-8. 10.1038/oby.2004.68.

Medina EA, Horn WF, Keim NL, Havel PJ, Benito P, Kelley DS, Nelson GJ, Erickson KL: Conjugated linoleic acid supplementation in humans: effects on circulating leptin concentrations and appetite. Lipids. 2000, 35 (7): 783-8. 10.1007/s11745-000-0586-y.

Salas-Salvado J, Marquez-Sandoval F, Bullo M: Conjugated linoleic acid intake in humans: a systematic review focusing on its effect on body composition, glucose, and lipid metabolism. Crit Rev Food Sci Nutr. 2006, 46 (6): 479-88. 10.1080/10408390600723953.

Von Loeffelholz C: Influence of conjugated linoleic acid (CLA) supplementation on body composition and strength in bodybuilders. Jena (Thnr). 1999, 7: 238-43.

Wang Y, Jones PJ: Dietary conjugated linoleic acid and body composition. Am J Clin Nutr. 2004, 79 (6 Suppl): 1153S-8S.

Wang YW, Jones PJ: Conjugated linoleic acid and obesity control: efficacy and mechanisms. Int J Obes Relat Metab Disord. 2004, 28 (8): 941-55. 10.1038/sj.ijo.0802641.

Zambell KL, Keim NL, Van Loan MD, Gale B, Benito P, Kelley DS, Nelson GJ: Conjugated linoleic acid supplementation in humans: effects on body composition and energy expenditure. Lipids. 2000, 35 (7): 777-82. 10.1007/s11745-000-0585-z.

Sneddon AA, Tsofliou F, Fyfe CL, Matheson I, Jackson DM, Horgan G, Winzell MS, Wahle KW, Ahren B, Williams LM: Effect of a conjugated linoleic acid and omega-3 fatty acid mixture on body composition and adiponectin. Obesity (Silver Spring). 2008, 16 (5): 1019-24. 10.1038/oby.2008.41.

Shigematsu N, Asano R, Shimosaka M, Okazaki M: Effect of administration with the extract of Gymnema sylvestre R. Br leaves on lipid metabolism in rats. Biol Pharm Bull. 2001, 24 (6): 713-7. 10.1248/bpb.24.713.

Shigematsu N, Asano R, Shimosaka M, Okazaki M: Effect of long term-administration with Gymnema sylvestre R. BR on plasma and liver lipid in rats. Biol Pharm Bull. 2001, 24 (6): 643-9. 10.1248/bpb.24.643.

Luo H, Kashiwagi A, Shibahara T, Yamada K: Decreased bodyweight without rebound and regulated lipoprotein metabolism by gymnemate in genetic multifactor syndrome animal. Mol Cell Biochem. 2007, 299 (1-2): 93-8. 10.1007/s11010-005-9049-7.

Preuss HG, Rao CV, Garis R, Bramble JD, Ohia SE, Bagchi M, Bagchi D: An overview of the safety and efficacy of a novel, natural(-)-hydroxycitric acid extract (HCA-SX) for weight management. J Med. 2004, 35 (1-6): 33-48.

Garcia Neto M, Pesti GM, Bakalli RI: Influence of dietary protein level on the broiler chicken's response to methionine and betaine supplements. Poult Sci. 2000, 79 (10): 1478-84.

Schwab U, Torronen A, Toppinen L, Alfthan G, Saarinen M, Aro A, Uusitupa M: Betaine supplementation decreases plasma homocysteine concentrations but does not affect body weight, body composition, or resting energy expenditure in human subjects. Am J Clin Nutr. 2002, 76 (5): 961-7.

Hoffman JR, Ratamess NA, Kang J, Rashti SL, Faigenbaum AD: Effect of betaine supplementation on power performance and fatigue. J Int Soc Sports Nutr. 2009, 6: 7-10.1186/1550-2783-6-7.

Ammon HP, Muller AB: Forskolin: from an ayurvedic remedy to a modern agent. Planta Med. 1985, 473-7. 10.1055/s-2007-969566. 6

Ammon HP, Muller AB: Effect of forskolin on islet cyclic AMP, insulin secretion, blood glucose and intravenous glucose tolerance in rats. Naunyn Schmiedebergs Arch Pharmacol. 1984, 326 (4): 364-7. 10.1007/BF00501444.

de Souza NJ, Dohadwalla AN, Reden J: Forskolin: a labdane diterpenoid with antihypertensive, positive inotropic, platelet aggregation inhibitory, and adenylate cyclase activating properties. Med Res Rev. 1983, 3 (2): 201-19. 10.1002/med.2610030205.

Litosch I, Hudson TH, Mills I, Li SY, Fain JN: Forskolin as an activator of cyclic AMP accumulation and lipolysis in rat adipocytes. Mol Pharmacol. 1982, 22 (1): 109-15.

Litosch I, Saito Y, Fain JN: Forskolin as an activator of cyclic AMP accumulation and secretion in blowfly salivary glands. Biochem J. 1982, 204 (1): 147-51.

Seamon KB, Padgett W, Daly JW: Forskolin: unique diterpene activator of adenylate cyclase in membranes and in intact cells. Proc Natl Acad Sci USA. 1981, 78 (6): 3363-7. 10.1073/pnas.78.6.3363.

Henderson S, Magu B, Rasmussen C, Lancaster S, Kerksick C, Smith P, Melton C, Cowan P, Greenwood M, Earnest C, Almada A, Milnor P, Magrans T, Bowden R, Ounpraseuth S, Thomas A, Kreider RB: Effects of coleus forskohlii supplementation on body composition and hematological profiles in mildly overweight women. J Int Soc Sports Nutr. 2005, 2: 54-62. 10.1186/1550-2783-2-2-54.

Godard MP, Johnson BA, Richmond SR: Body composition and hormonal adaptations associated with forskolin consumption in overweight and obese men. Obes Res. 2005, 13 (8): 1335-43. 10.1038/oby.2005.162.

Kreider RB, Henderson S, Magu B, Rasmussen C, Lancaster S, Kerksick C, Smith P, Melton C, Cowan P, Greenwood M, Earnest C, Almada A, Milnor P: Effects of coleus forskohlii supplementation on body composition and markers of health in sedentary overweight females. FASEB J. 2002, LB59-

Ebeling P, Koivisto VA: Physiological importance of dehydroepiandrosterone. Lancet. 1994, 343 (8911): 1479-81. 10.1016/S0140-6736(94)92587-9.

Denti L, Pasolini G, Sanfelici L, Ablondi F, Freddi M, Benedetti R, Valenti G: Effects of aging on dehydroepiandrosterone sulfate in relation to fasting insulin levels and body composition assessed by bioimpedance analysis. Metabolism. 1997, 46 (7): 826-32. 10.1016/S0026-0495(97)90130-X.

De Pergola G, Zamboni M, Sciaraffia M, Turcato E, Pannacciulli N, Armellini F, Giorgino F, Perrini S, Bosello O, Giorgino R: Body fat accumulation is possibly responsible for lower dehydroepiandrosterone circulating levels in premenopausal obese women. Int J Obes Relat Metab Disord. 1996, 20 (12): 1105-10.

Nestler JE, Barlascini CO, Clore JN, Blackard WG: Dehydroepiandrosterone reduces serum low density lipoprotein levels and body fat but does not alter insulin sensitivity in normal men. J Clin Endocrinol Metab. 1988, 66 (1): 57-61. 10.1210/jcem-66-1-57.

Vogiatzi MG, Boeck MA, Vlachopapadopoulou E, el-Rashid R, New MI: Dehydroepiandrosterone in morbidly obese adolescents: effects on weight, body composition, lipids, and insulin resistance. Metabolism. 1996, 45 (8): 1011-5. 10.1016/S0026-0495(96)90272-3.

von Muhlen D, Laughlin GA, Kritz-Silverstein D, Bergstrom J, Bettencourt R: Effect of dehydroepiandrosterone supplementation on bone mineral density, bone markers, and body composition in older adults: the DAWN trial. Osteoporos Int. 2008, 19 (5): 699-707. 10.1007/s00198-007-0520-z.

Kalman DS, Colker CM, Swain MA, Torina GC, Shi Q: A randomized double-blind, placebo-controlled study of 3-acetyl-7-oxo-dehydroepiandrosterone in healthy overweight adults. Curr Thera. 2000, 61: 435-42. 10.1016/S0011-393X(00)80026-0.

Zenk JL, Frestedt JL, Kuskowski MA: HUM5007, a novel combination of thermogenic compounds, and 3-acetyl-7-oxo-dehydroepiandrosterone: each increases the resting metabolic rate of overweight adults. J Nutr Biochem. 2007, 18 (9): 629-34. 10.1016/j.jnutbio.2006.11.008.

Zenk JL, Leikam SA, Kassen LJ, Kuskowski MA: Effect of lean system 7 on metabolic rate and body composition. Nutrition. 2005, 21 (2): 179-85. 10.1016/j.nut.2004.05.025.

Stanko RT, Arch JE: Inhibition of regain in body weight and fat with addition of 3-carbon compounds to the diet with hyperenergetic refeeding after weight reduction. Int J Obes Relat Metab Disord. 1996, 20 (10): 925-30.

Stanko RT, Tietze DL, Arch JE: Body composition, energy utilization, and nitrogen metabolism with a severely restricted diet supplemented with dihydroxyacetone and pyruvate. Am J Clin Nutr. 1992, 55 (4): 771-6.

Stanko RT, Reynolds HR, Hoyson R, Janosky JE, Wolf R: Pyruvate supplementation of a low-cholesterol, low-fat diet: effects on plasma lipid concentrations and body composition in hyperlipidemic patients. Am J Clin Nutr. 1994, 59 (2): 423-7.

Kalman D, Colker CM, Wilets I, Roufs JB, Antonio J: The effects of pyruvate supplementation on body composition in overweight individuals. Nutrition. 1999, 15 (5): 337-40. 10.1016/S0899-9007(99)00034-9.

Stone MH, Sanborn K, Smith LL, O'Bryant HS, Hoke T, Utter AC, Johnson RL, Boros R, Hruby J, Pierce KC, Stone ME, Garner B: Effects of in-season (5 weeks) creatine and pyruvate supplementation on anaerobic performance and body composition in American football players. Int J Sport Nutr. 1999, 9 (2): 146-65.

Koh-Banerjee PK, Ferreira MP, Greenwood M, Bowden RG, Cowan PN, Almada AL, Kreider RB: Effects of calcium pyruvate supplementation during training on body composition, exercise capacity, and metabolic responses to exercise. Nutrition. 2005, 21 (3): 312-9. 10.1016/j.nut.2004.06.026.

Gallaher DD, Gallaher CM, Mahrt GJ, Carr TP, Hollingshead CH, Hesslink R, Wise J: A glucomannan and chitosan fiber supplement decreases plasma cholesterol and increases cholesterol excretion in overweight normocholesterolemic humans. J Am Coll Nutr. 2002, 21 (5): 428-33.

Gallaher CM, Munion J, Hesslink R, Wise J, Gallaher DD: Cholesterol reduction by glucomannan and chitosan is mediated by changes in cholesterol absorption and bile acid and fat excretion in rats. J Nutr. 2000, 130 (11): 2753-9.

Chiang MT, Yao HT, Chen HC: Effect of dietary chitosans with different viscosity on plasma lipids and lipid peroxidation in rats fed on a diet enriched with cholesterol. Biosci Biotechnol Biochem. 2000, 64 (5): 965-71. 10.1271/bbb.64.965.

Tai TS, Sheu WH, Lee WJ, Yao HT, Chiang MT: Effect of chitosan on plasma lipoprotein concentrations in type 2 diabetic subjects with hypercholesterolemia. Diabetes Care. 2000, 23 (11): 1703-4. 10.2337/diacare.23.11.1703a.

Wuolijoki E, Hirvela T, Ylitalo P: Decrease in serum LDL cholesterol with microcrystalline chitosan. Methods Find Exp Clin Pharmacol. 1999, 21 (5): 357-61. 10.1358/mf.1999.21.5.793477.

Gades MD, Stern JS: Chitosan supplementation and fecal fat excretion in men. Obes Res. 2003, 11 (5): 683-8. 10.1038/oby.2003.97.

Guerciolini R, Radu-Radulescu L, Boldrin M, Dallas J, Moore R: Comparative evaluation of fecal fat excretion induced by orlistat and chitosan. Obes Res. 2001, 9 (6): 364-7. 10.1038/oby.2001.47.

Gades MD, Stern JS: Chitosan supplementation and fat absorption in men and women. J Am Diet Assoc. 2005, 105 (1): 72-7. 10.1016/j.jada.2004.10.004.

Pittler MH, Abbot NC, Harkness EF, Ernst E: Randomized, double-blind trial of chitosan for body weight reduction. Eur J Clin Nutr. 1999, 53 (5): 379-81. 10.1038/sj.ejcn.1600733.

Ho SC, Tai ES, Eng PH, Tan CE, Fok AC: In the absence of dietary surveillance, chitosan does not reduce plasma lipids or obesity in hypercholesterolaemic obese Asian subjects. Singapore Med J. 2001, 42 (1): 006-10.

Vincent J: The potential value and toxicity of chromium picolinate as a nutritional supplement, weight loss agent and muscle development agent. Sports Med. 2003, 33 (3): 213-30. 10.2165/00007256-200333030-00004.

Lukaski HC, Siders WA, Penland JG: Chromium picolinate supplementation in women: effects on body weight, composition, and iron status. Nutrition. 2007, 23 (3): 187-95. 10.1016/j.nut.2006.12.001.

Jena BS, Jayaprakasha GK, Singh RP, Sakariah KK: Chemistry and biochemistry of (-)-hydroxycitric acid from Garcinia. J Agric Food Chem. 2002, 50 (1): 10-22. 10.1021/jf010753k.

Ishihara K, Oyaizu S, Onuki K, Lim K, Fushiki T: Chronic (-)-hydroxycitrate administration spares carbohydrate utilization and promotes lipid oxidation during exercise in mice. J Nutr. 2000, 130 (12): 2990-5.

Kriketos AD, Thompson HR, Greene H, Hill JO: (-)-Hydroxycitric acid does not affect energy expenditure and substrate oxidation in adult males in a post-absorptive state. Int J Obes Relat Metab Disord. 1999, 23 (8): 867-73. 10.1038/sj.ijo.0800965.

Heymsfield SB, Allison DB, Vasselli JR, Pietrobelli A, Greenfield D, Nunez C: Garcinia cambogia (hydroxycitric acid) as a potential antiobesity agent: a randomized controlled trial. Jama. 1998, 280 (18): 1596-600. 10.1001/jama.280.18.1596.

Mattes RD, Bormann L: Effects of (-)-hydroxycitric acid on appetitive variables. Physiol Behav. 2000, 71 (1-2): 87-94. 10.1016/S0031-9384(00)00321-8.

Kraemer WJ, Volek JS, Dunn-Lewis C: L-carnitine supplementation: influence upon physiological function. Curr Sports Med Rep. 2008, 7 (4): 218-23.

Smith WA, Fry AC, Tschume LC, Bloomer RJ: Effect of glycine propionyl-L-carnitine on aerobic and anaerobic exercise performance. Int J Sport Nutr Exerc Metab. 2008, 18 (1): 19-36.

Brass EP: Supplemental carnitine and exercise. Am J Clin Nutr. 2000, 72 (2 Suppl): 618S-23S.

Villani RG, Gannon J, Self M, Rich PA: L-Carnitine supplementation combined with aerobic training does not promote weight loss in moderately obese women. Int J Sport Nutr Exerc Metab. 2000, 10 (2): 199-207.

Bloomer RJ, Smith WA: Oxidative stress in response to aerobic and anaerobic power testing: influence of exercise training and carnitine supplementation. Res Sports Med. 2009, 17 (1): 1-16. 10.1080/15438620802678289.

Volek JS, Kraemer WJ, Rubin MR, Gomez AL, Ratamess NA, Gaynor P: L-Carnitine L-tartrate supplementation favorably affects markers of recovery from exercise stress. Am J Physiol Endocrinol Metab. 2002, 282 (2): E474-82.

Kaciuba-Uscilko H, Nazar K, Chwalbinska-Moneta J, Ziemba A, Kruk B, Szczepanik J, Titow-Stupnicka E, Bicz B: Effect of phosphate supplementation on metabolic and neuroendocrine responses to exercise and oral glucose load in obese women during weight reduction. J Physiol Pharmacol. 1993, 44 (4): 425-40.

Nazar K, Kaciuba-Uscilko H, Szczepanik J, Zemba AW, Kruk B, Chwalbinska-Moneta J, Titow-Stupnicka E, Bicz B, Krotkiewski M: Phosphate supplementation prevents a decrease of triiodothyronine and increases resting metabolic rate during low energy diet. J Physiol Pharmacol. 1996, 47 (2): 373-83.

Grases F, Llompart I, Conte A, Coll R, March JG: Glycosaminoglycans and oxalocalcic urolithiasis. Nephron. 1994, 68 (4): 449-53. 10.1159/000188306.

Grases F, Melero G, Costa-Bauza A, Prieto R, March JG: Urolithiasis and phytotherapy. Int Urol Nephrol. 1994, 26 (5): 507-11. 10.1007/BF02767650.

Dolan RL, Crosby EC, Leutkemeir MJ, Barton RG, Askew EW: The effects of diuretics on resting metabolic rate and subsequent shifts in respiratory exchange ratios. Med Sci Sports Exerc. 2001, 33: S163-10.1097/00005768-200105001-00921.

Crosby EC, Dolan RL, Benson JE, Leutkemeir MJ, Barton RG, Askew EW: Herbal diuretic induced dehydration and resting metabolic rate. Med Sci Sports Exerc. 2001, 33: S163-10.1097/00005768-200105001-00920.

Von Duvillard SP, Braun WA, Markofski M, Beneke R, Leithauser R: Fluids and hydration in prolonged endurance performance. Nutrition. 2004, 20 (7-8): 651-6. 10.1016/j.nut.2004.04.011.

von Duvillard SP, Arciero PJ, Tietjen-Smith T, Alford K: Sports drinks, exercise training, and competition. Curr Sports Med Rep. 2008, 7 (4): 202-8.

Winnick JJ, Davis JM, Welsh RS, Carmichael MD, Murphy EA, Blackmon JA: Carbohydrate feedings during team sport exercise preserve physical and CNS function. Med Sci Sports Exerc. 2005, 37 (2): 306-15. 10.1249/01.MSS.0000152803.35130.A4.

Kendall RW, Jacquemin G, Frost R, Burns SP: Creatine supplementation for weak muscles in persons with chronic tetraplegia: a randomized double-blind placebo-controlled crossover trial. J Spinal Cord Med. 2005, 28 (3): 208-13.

Kendall KL, Smith AE, Graef JL, Fukuda DH, Moon JR, Beck TW, Cramer JT, Stout JR: Effects of four weeks of high-intensity interval training and creatine supplementation on critical power and anaerobic working capacity in college-aged men. J Strength Cond Res. 2009, 23 (6): 1663-9.

Kreider RB, Ferreira M, Wilson M, Grindstaff P, Plisk S, Reinardy J, Cantler E, Almada AL: Effects of creatine supplementation on body composition, strength, and sprint performance. Med Sci Sports Exerc. 1998, 30 (1): 73-82.

Derave W, Op'T Eijinde B, Richter EA, Hespel P: Combined creatine and protein supplementation improves glucose tolerance and muscle glycogen accumulation in humans. Abstracts of 6th Internationl Conference on Guanidino Compounds in Biology and Medicine. 2001

Nelson AG, Arnall DA, Kokkonen J, Day R, Evans J: Muscle glycogen supercompensation is enhanced by prior creatine supplementation. Med Sci Sports Exerc. 2001, 33 (7): 1096-100.

Op 't Eijnde B, Richter EA, Henquin JC, Kiens B, Hespel P: Effect of creatine supplementation on creatine and glycogen content in rat skeletal muscle. Acta Physiol Scand. 2001, 171 (2): 169-76. 10.1046/j.1365-201x.2001.00786.x.

Chwalbinska-Moneta J: Effect of creatine supplementation on aerobic performance and anaerobic capacity in elite rowers in the course of endurance training. Int J Sport Nutr Exerc Metab. 2003, 13 (2): 173-83.

Green AL, Hultman E, Macdonald IA, Sewell DA, Greenhaff P: Carbohydrate feeding augments skeletal muscle creatine accumulation during creatine supplementation in humans. Am J Physiol. 1996, 271: E821-E6.

Nelson AG, Day R, Glickman-Weiss EL, Hegsted M, Kokkonen J, Sampson B: Creatine supplementation alters the response to a graded cycle ergometer test. Eur J Appl Physiol. 2000, 83 (1): 89-94. 10.1007/s004210000244.

Nelson AG, Day R, Glickman-Weiss EL, Hegsted M, Sampson B: Creatine supplementation raises anaerobic threshold. FASEB Journal. 1997, 11: A589-

Kreider RB, Miller GW, Williams MH, Somma CT, Nasser TA: Effects of phosphate loading on oxygen uptake, ventilatory anaerobic threshold, and run performance. Med Sci Sports Exerc. 1990, 22 (2): 250-6.

Cade R, Conte M, Zauner C, Mars D, Peterson J, Lunne D, Hommen N, Packer D: Effects of phosphate loading on 2,3 diphosphoglycerate and maximal oxygen uptake. Med Sci Sports Exerc. 1984, 16: 263-8.

Kreider RB, Miller GW, Schenck D, Cortes CW, Miriel V, Somma CT, Rowland P, Turner C, Hill D: Effects of phosphate loading on metabolic and myocardial responses to maximal and endurance exercise. Int J Sport Nutr. 1992, 2 (1): 20-47.

Stewart I, McNaughton L, Davies P, Tristram S: Phosphate loading and the effects of VO2max in trained cyclists. Res Quart. 1990, 61: 80-4.

Folland JP, Stern R, Brickley G: Sodium phosphate loading improves laboratory cycling time-trial performance in trained cyclists. J Sci Med Sport. 2008, 11 (5): 464-8. 10.1016/j.jsams.2007.04.004.

McNaughton L, Backx K, Palmer G, Strange N: Effects of chronic bicarbonate ingestion on the performance of high-intensity work. Eur J Appl Physiol Occup Physiol. 1999, 80 (4): 333-6. 10.1007/s004210050600.

Applegate E: Effective nutritional ergogenic aids. Int J Sport Nutr. 1999, 9 (2): 229-39.

Kronfeld DS, Ferrante PL, Grandjean D: Optimal nutrition for athletic performance, with emphasis on fat adaptation in dogs and horses. J Nutr. 1994, 124 (12 Suppl): 2745S-53S.

Kraemer WJ, Gordon SE, Lynch JM, Pop ME, Clark KL: Effects of multibuffer supplementation on acid-base balance and 2,3-diphosphoglycerate following repetitive anaerobic exercise. Int J Sport Nutr. 1995, 5 (4): 300-14.

Matson LG, Tran ZV: Effects of sodium bicarbonate ingestion on anaerobic performance: a meta-analytic review. Int J Sport Nutr. 1993, 3 (1): 2-28.

Lindh AM, Peyrebrune MC, Ingham SA, Bailey DM, Folland JP: Sodium bicarbonate improves swimming performance. Int J Sports Med. 2008, 29 (6): 519-23. 10.1055/s-2007-989228.

Wiles JD, Coleman D, Tegerdine M, Swaine IL: The effects of caffeine ingestion on performance time, speed and power during a laboratory-based 1 km cycling time-trial. J Sports Sci. 2006, 24 (11): 1165-71. 10.1080/02640410500457687.

Ivy JL, Kammer L, Ding Z, Wang B, Bernard JR, Liao YH, Hwang J: Improved cycling time-trial performance after ingestion of a caffeine energy drink. Int J Sport Nutr Exerc Metab. 2009, 19 (1): 61-78.

McNaughton LR, Lovell RJ, Siegler J, Midgley AW, Moore L, Bentley DJ: The effects of caffeine ingestion on time trial cycling performance. Int J Sports Physiol Perform. 2008, 3 (2): 157-63.

Graham TE: Caffeine and exercise: metabolism, endurance and performance. Sports Med. 2001, 31 (11): 785-807. 10.2165/00007256-200131110-00002.

Carr A, Dawson B, Schneiker K, Goodman C, Lay B: Effect of caffeine supplementation on repeated sprint running performance. J Sports Med Phys Fitness. 2008, 48 (4): 472-8.

Glaister M, Howatson G, Abraham CS, Lockey RA, Goodwin JE, Foley P, McInnes G: Caffeine supplementation and multiple sprint running performance. Med Sci Sports Exerc. 2008, 40 (10): 1835-40. 10.1249/MSS.0b013e31817a8ad2.

Tarnopolsky MA, Atkinson SA, MacDougall JD, Sale DG, Sutton JR: Physiological responses to caffeine during endurance running in habitual caffeine users. Med Sci Sports Exerc. 1989, 21 (4): 418-24.

Armstrong LE: Caffeine, body fluid-electrolyte balance, and exercise performance. Int J Sport Nutr Exerc Metab. 2002, 12 (2): 189-206.

Falk B, Burstein R, Rosenblum J, Shapiro Y, Zylber-Katz E, Bashan N: Effects of caffeine ingestion on body fluid balance and thermoregulation during exercise. Can J Physiol Pharmacol. 1990, 68 (7): 889-92.

Harris R, Dunnett M, Greenhaf P: Carnosine and Taurine contents in individual fibres of human vastus lateralis muscle. J Sport Sci. 1998, 16: 639-43. 10.1080/026404198366443.

Harris RC, Tallon MJ, Dunnett M, Boobis L, Coakley J, Kim HJ, Fallowfield JL, Hill CA, Sale C, Wise JA: The absorption of orally supplied beta-alanine and its effect on muscle carnosine synthesis in human vastus lateralis. Amino Acids. 2006, 30 (3): 279-89. 10.1007/s00726-006-0299-9.

Stout JR, Cramer JT, Mielke M, O'Kroy J, Torok DJ, Zoeller RF: Effects of twenty-eight days of beta-alanine and creatine monohydrate supplementation on the physical working capacity at neuromuscular fatigue threshold. J Strength Cond Res. 2006, 20 (4): 928-31. 10.1519/R-19655.1.

Hill CA, Harris RC, Kim HJ, Harris BD, Sale C, Boobis LH, Kim CK, Wise JA: Influence of beta-alanine supplementation on skeletal muscle carnosine concentrations and high intensity cycling capacity. Amino Acids. 2007, 32 (2): 225-33. 10.1007/s00726-006-0364-4.

Hoffman J, Ratamess NA, Ross R, Kang J, Magrelli J, Neese K, Faigenbaum AD, Wise JA: beta-Alanine and the Hormonal Response to Exercise. Int J Sports Med. 2008, 29 (12): 952-8. 10.1055/s-2008-1038678.

Smith AE, Walter AA, Graef JL, Kendall KL, Moon JR, Lockwood CM, Fakuda DH, Beck TW, Cramer JT, Stout JR: Effects of beta-alanine supplementation and high-intensity interval training on endurance performance and body composition in men; a double-blind trial. J Int Soc Sports Nutr. 2009, 6 (1): 5. 10.1186/1550-2783-6-5.

Derave W, Ozdemir MS, Harris RC, Pottier A, Reyngoudt H, Koppo K, Wise JA, Achten E: beta-Alanine supplementation augments muscle carnosine content and attenuates fatigue during repeated isokinetic contraction bouts in trained sprinters. J Appl Physiol. 2007, 103 (5): 1736-43. 10.1152/japplphysiol.00397.2007.

Hoffman JR, Ratamess NA, Faigenbaum AD, Ross R, Kang J, Stout JR, Wise JA: Short-duration beta-alanine supplementation increases training volume and reduces subjective feelings of fatigue in college football players. Nutr Res. 2008, 28 (1): 31-5. 10.1016/j.nutres.2007.11.004.

Hoffman J, Ratamess N, Kang J, Mangine G, Faigenbaum A, Stout J: Effect of creatine and beta-alanine supplementation on performance and endocrine responses in strength/power athletes. Int J Sport Nutr Exerc Metab. 2006, 16 (4): 430-46.

Kendrick IP, Harris RC, Kim HJ, Kim CK, Dang VH, Lam TQ, Bui TT, Smith M, Wise JA: The effects of 10 weeks of resistance training combined with beta-alanine supplementation on whole body strength, force production, muscular endurance and body composition. Amino Acids. 2008, 34 (4): 547-54. 10.1007/s00726-007-0008-3.

Tarnopolsky MA, Parise G, Yardley NJ, Ballantyne CS, Olatinji S, Phillips SM: Creatine-dextrose and protein-dextrose induce similar strength gains during training. Med Sci Sports Exerc. 2001, 33 (12): 2044-52. 10.1097/00005768-200112000-00011.

Kreider RB, Earnest CP, Lundberg J, Rasmussen C, Greenwood M, Cowan P, Almada AL: Effects of ingesting protein with various forms of carbohydrate following resistance-exercise on substrate availability and markers of anabolism, catabolism, and immunity. J Int Soc Sports Nutr. 2007, 4: 18-10.1186/1550-2783-4-18.

Cribb PJ, Hayes A: Effects of supplement timing and resistance exercise on skeletal muscle hypertrophy. Med Sci Sports Exerc. 2006, 38 (11): 1918-25. 10.1249/01.mss.0000233790.08788.3e.

Kerksick CM, Rasmussen CJ, Lancaster SL, Magu B, Smith P, Melton C, Greenwood M, Almada AL, Earnest CP, Kreider RB: The effects of protein and amino acid supplementation on performance and training adaptations during ten weeks of resistance training. J Strength Cond Res. 2006, 20 (3): 643-53. 10.1519/R-17695.1.

Tipton KD, Borsheim E, Wolf SE, Sanford AP, Wolfe RR: Acute response of net muscle protein balance reflects 24-h balance after exercise and amino acid ingestion. Am J Physiol Endocrinol Metab. 2003, 284: E76-E89.

Hoffman JR, Cooper J, Wendell M, Im J, Kang J: Effects of beta-hydroxy beta-methylbutyrate on power performance and indices of muscle damage and stress during high-intensity training. J Strength Cond Res. 2004, 18 (4): 747-52. 10.1519/13973.1.

Thomson JS, Watson PE, Rowlands DS: Effects of nine weeks of beta-hydroxy-beta-methylbutyrate supplementation on strength and body composition in resistance trained men. J Strength Cond Res. 2009, 23 (3): 827-35.

Wagner DR: Hyperhydrating with glycerol: implications for athletic performance. J Am Diet Assoc. 1999, 99 (2): 207-12. 10.1016/S0002-8223(99)00049-8.

Inder WJ, Swanney MP, Donald RA, Prickett TC, Hellemans J: The effect of glycerol and desmopressin on exercise performance and hydration in triathletes. Med Sci Sports Exerc. 1998, 30 (8): 1263-9. 10.1097/00005768-199808000-00013.

Montner P, Stark DM, Riedesel ML, Murata G, Robergs R, Timms M, Chick TW: Pre-exercise glycerol hydration improves cycling endurance time. Int J Sports Med. 1996, 17 (1): 27-33. 10.1055/s-2007-972804.

Boulay MR, Song TM, Serresse O, Theriault G, Simoneau JA, Bouchard C: Changes in plasma electrolytes and muscle substrates during short-term maximal exercise in humans. Can J Appl Physiol. 1995, 20 (1): 89-101.

Tikuisis P, Ducharme MB, Moroz D, Jacobs I: Physiological responses of exercised-fatigued individuals exposed to wet-cold conditions. J Appl Physiol. 1999, 86 (4): 1319-28.

Jimenez C, Melin B, Koulmann N, Allevard AM, Launay JC, Savourey G: Plasma volume changes during and after acute variations of body hydration level in humans. Eur J Appl Physiol Occup Physiol. 1999, 80 (1): 1-8. 10.1007/s004210050550.

Magal M, Webster MJ, Sistrunk LE, Whitehead MT, Evans RK, Boyd JC: Comparison of glycerol and water hydration regimens on tennis-related performance. Med Sci Sports Exerc. 2003, 35 (1): 150-6. 10.1097/00005768-200301000-00023.

Kavouras SA, Armstrong LE, Maresh CM, Casa DJ, Herrera-Soto JA, Scheett TP, Stoppani J, Mack GW, Kraemer WJ: Rehydration with glycerol: endocrine, cardiovascular, and thermoregulatory responses during exercise in the heat. J Appl Physiol. 2006, 100 (2): 442-50. 10.1152/japplphysiol.00187.2005.

Jeukendrup AE, Thielen JJ, Wagenmakers AJ, Brouns F, Saris WH: Effect of medium-chain triacylglycerol and carbohydrate ingestion during exercise on substrate utilization and subsequent cycling performance. Am J Clin Nutr. 1998, 67 (3): 397-404.

Goedecke JH, Elmer-English R, Dennis SC, Schloss I, Noakes TD, Lambert EV: Effects of medium-chain triaclyglycerol ingested with carbohydrate on metabolism and exercise performance. Int J Sport Nutr. 1999, 9 (1): 35-47.

Calabrese C, Myer S, Munson S, Turet P, Birdsall TC: A cross-over study of the effect of a single oral feeding of medium chain triglyceride oil vs. canola oil on post-ingestion plasma triglyceride levels in healthy men. Altern Med Rev. 1999, 4 (1): 23-8.

Angus DJ, Hargreaves M, Dancey J, Febbraio MA: Effect of carbohydrate or carbohydrate plus medium-chain triglyceride ingestion on cycling time trial performance. J Appl Physiol. 2000, 88 (1): 113-9.

Van Zyl CG, Lambert EV, Hawley JA, Noakes TD, Dennis SC: Effects of medium-chain triglyceride ingestion on fuel metabolism and cycling performance. J Appl Physiol. 1996, 80 (6): 2217-25.

Misell LM, Lagomarcino ND, Schuster V, Kern M: Chronic medium-chain triacylglycerol consumption and endurance performance in trained runners. J Sports Med Phys Fitness. 2001, 41 (2): 210-5.

Goedecke JH, Clark VR, Noakes TD, Lambert EV: The effects of medium-chain triacylglycerol and carbohydrate ingestion on ultra-endurance exercise performance. Int J Sport Nutr Exerc Metab. 2005, 15 (1): 15-27.

Burke LM, Kiens B, Ivy JL: Carbohydrates and fat for training and recovery. J Sports Sci. 2004, 22 (1): 15-30. 10.1080/0264041031000140527.

Thorburn MS, Vistisen B, Thorp RM, Rockell MJ, Jeukendrup AE, Xu X, Rowlands DS: Attenuated gastric distress but no benefit to performance with adaptation to octanoate-rich esterified oils in well-trained male cyclists. J Appl Physiol. 2006, 101 (6): 1733-43. 10.1152/japplphysiol.00393.2006.

Nosaka N, Suzuki Y, Nagatoishi A, Kasai M, Wu J, Taguchi M: Effect of ingestion of medium-chain triacylglycerols on moderate- and high-intensity exercise in recreational athletes. J Nutr Sci Vitaminol (Tokyo). 2009, 55 (2): 120-5. 10.3177/jnsv.55.120.

Tullson PC, Terjung RL: Adenine nucleotide synthesis in exercising and endurance-trained skeletal muscle. Am J Physiol. 1991, 261 (2 Pt 1): C342-7.

Gross M, Kormann B, Zollner N: Ribose administration during exercise: effects on substrates and products of energy metabolism in healthy subjects and a patient with myoadenylate deaminase deficiency. Klin Wochenschr. 1991, 69 (4): 151-5. 10.1007/BF01665856.

Wagner DR, Gresser U, Kamilli I, Gross M, Zollner N: Effects of oral ribose on muscle metabolism during bicycle ergometer in patients with AMP-deaminase-deficiency. Adv Exp Med Biol. 1991, 383-5.

Pliml W, von Arnim T, Stablein A, Hofmann H, Zimmer HG, Erdmann E: Effects of ribose on exercise-induced ischaemia in stable coronary artery disease. Lancet. 1992, 340 (8818): 507-10. 10.1016/0140-6736(92)91709-H.

Pauly DF, Pepine CJ: D-Ribose as a supplement for cardiac energy metabolism. J Cardiovasc Pharmacol Ther. 2000, 5 (4): 249-58. 10.1054/JCPT.2000.18011.

Op 't Eijnde B, Van Leemputte M, Brouns F, Vusse Van Der GJ, Labarque V, Ramaekers M, Van Schuylenberg R, Verbessem P, Wijnen H, Hespel P: No effects of oral ribose supplementation on repeated maximal exercise and de novo ATP resynthesis. J Appl Physiol. 2001, 91 (5): 2275-81.

Berardi JM, Ziegenfuss TN: Effects of ribose supplementation on repeated sprint performance in men. J Strength Cond Res. 2003, 17 (1): 47-52. 10.1519/1533-4287(2003)017<0047:EORSOR>2.0.CO;2.

Kreider RB, Melton C, Greenwood M, Rasmussen C, Lundberg J, Earnest C, Almada A: Effects of oral D-ribose supplementation on anaerobic capacity and selected metabolic markers in healthy males. Int J Sport Nutr Exerc Metab. 2003, 13 (1): 76-86.

Dunne L, Worley S, Macknin M: Ribose versus dextrose supplementation, association with rowing performance: a double-blind study. Clin J Sport Med. 2006, 16 (1): 68-71. 10.1097/01.jsm.0000180022.44889.94.

Kerksick C, Rasmussen C, Bowden R, Leutholtz B, Harvey T, Earnest C, Greenwood M, Almada A, Kreider R: Effects of ribose supplementation prior to and during intense exercise on anaerobic capacity and metabolic markers. Int J Sport Nutr Exerc Metab. 2005, 15 (6): 653-64.

Hargreaves M, McKenna MJ, Jenkins DG, Warmington SA, Li JL, Snow RJ, Febbraio MA: Muscle metabolites and performance during high-intensity, intermittent exercise. J Appl Physiol. 1998, 84 (5): 1687-91.

Starling RD, Trappe TA, Short KR, Sheffield-Moore M, Jozsi AC, Fink WJ, Costill DL: Effect of inosine supplementation on aerobic and anaerobic cycling performance. Med Sci Sports Exerc. 1996, 28 (9): 1193-8.

Williams MH, Kreider RB, Hunter DW, Somma CT, Shall LM, Woodhouse ML, Rokitski L: Effect of inosine supplementation on 3-mile treadmill run performance and VO2 peak. Med Sci Sports Exerc. 1990, 22 (4): 517-22.

McNaughton L, Dalton B, Tarr J: Inosine supplementation has no effect on aerobic or anaerobic cycling performance. Int J Sport Nutr. 1999, 9 (4): 333-44.

Braham R, Dawson B, Goodman C: The effect of glucosamine supplementation on people experiencing regular knee pain. Br J Sports Med. 2003, 37 (1): 45-9. 10.1136/bjsm.37.1.45. discussion 9

Vad V, Hong HM, Zazzali M, Agi N, Basrai D: Exercise recommendations in athletes with early osteoarthritis of the knee. Sports Med. 2002, 32 (11): 729-39. 10.2165/00007256-200232110-00004.

Nieman DC: Exercise immunology: nutritional countermeasures. Can J Appl Physiol. 2001, 26 (Suppl): S45-55.

Gleeson M, Lancaster GI, Bishop NC: Nutritional strategies to minimise exercise-induced immunosuppression in athletes. Can J Appl Physiol. 2001, 26 (Suppl): S23-35.

Gleeson M, Bishop NC: Elite athlete immunology: importance of nutrition. Int J Sports Med. 2000, 21 (Suppl 1): S44-50. 10.1055/s-2000-1451.

Nieman DC, Pedersen BK: Exercise and immune function. Recent developments. Sports Med. 1999, 27 (2): 73-80. 10.2165/00007256-199927020-00001.

Lowery L, Berardi JM, Ziegenfuss Antioxidants: Sports Supplements. Edited by: Antonio J, Stout J. 2001, Baltimore, MD: Lippincott, Williams & Wilkins, 260-78.

Gomez AL, Volek JS, Ratamess NA, Rubin MR, Wickham RB, Mazzetti SA, Doan BK, Newton RU, Kraemer WJ: Creatine supplementation enhances body composition during short-term reisstance training overreaching. Journal of Strength and Conditioning Research. 2000, 14 (3):

French DN, Volek JS, Ratamess NA, Mazzetti SA, Rubin MR, Gomez AL, Wickham RB, Doan BK, McGuigan MR, Scheett TP, Newton RU, Dorofeyeva E, Kraemer WJ: The effects of creatine supplementation on resting serum hormonal concentrations during short-term resistance training overreaching. Med Sci Sports & Exerc. 2001, 33 (5): S203-10.1097/00005768-200105001-01142.

Mero A: Leucine supplementation and intensive training. Sports Med. 1999, 27 (6): 347-58. 10.2165/00007256-199927060-00001.

Harris WS, Appel LJ: New guidelines focus on fish, fish oil, omega-3 fatty acids. American Heart Association, 2002(November 11), [ http://www.americanheart.org/presenter.jhtml?identifier=3065754 ]

Williams MH: Vitamin supplementation and athletic performance. Int J Vitam Nutr Res Suppl. 1989, 30: 163-91.

Reid IR: Therapy of osteoporosis: calcium, vitamin D, and exercise. Am J Med Sci. 1996, 312 (6): 278-86. 10.1097/00000441-199612000-00006.

Goldfarb AH: Antioxidants: role of supplementation to prevent exercise-induced oxidative stress. Med Sci Sports Exerc. 1993, 25 (2): 232-6.

Goldfarb AH: Nutritional antioxidants as therapeutic and preventive modalities in exercise-induced muscle damage. Can J Appl Physiol. 1999, 24 (3): 249-66.

Appell HJ, Duarte JA, Soares JM: Supplementation of vitamin E may attenuate skeletal muscle immobilization atrophy. Int J Sports Med. 1997, 18 (3): 157-60. 10.1055/s-2007-972612.

Tiidus PM, Houston ME: Vitamin E status and response to exercise training. Sports Med. 1995, 20 (1): 12-23. 10.2165/00007256-199520010-00002.

Craciun AM, Wolf J, Knapen MH, Brouns F, Vermeer C: Improved bone metabolism in female elite athletes after vitamin K supplementation. Int J Sports Med. 1998, 19 (7): 479-84. 10.1055/s-2007-971948.

Fogelholm M, Ruokonen I, Laakso JT, Vuorimaa T, Himberg JJ: Lack of association between indices of vitamin B1, B2, and B6 status and exercise-induced blood lactate in young adults. Int J Sport Nutr. 1993, 3 (2): 165-76.

Garg R, Malinow M, Pettinger M, Upson B, Hunninghake D: Niacin treatment increases plasma homocyst(e)ine levels. Am Heart J. 1999, 138 (6 Pt 1): 1082-7. 10.1016/S0002-8703(99)70073-6.

Alaswad K, O'Keefe JH, Moe RM: Combination drug therapy for dyslipidemia. Curr Atheroscler Rep. 1999, 1 (1): 44-9. 10.1007/s11883-999-0049-z.

Murray R, Bartoli WP, Eddy DE, Horn MK: Physiological and performance responses to nicotinic-acid ingestion during exercise. Med Sci Sports Exerc. 1995, 27 (7): 1057-62. 10.1249/00005768-199507000-00015.

Bonke D: Influence of vitamin B1, B6, and B12 on the control of fine motoric movements. Bibl Nutr Dieta. 1986 (38): 104-9.

Bonke D, Nickel B: Improvement of fine motoric movement control by elevated dosages of vitamin B1, B6, and B12 in target shooting. Int J Vitam Nutr Res Suppl. 1989, 30: 198-204.

Van Dyke DC, Stumbo PJ, Mary JB, Niebyl JR: Folic acid and prevention of birth defects. Dev Med Child Neurol. 2002, 44 (6): 426-9. 10.1017/S0012162201002316.

Mattson MP, Kruman II, Duan W: Folic acid and homocysteine in age-related disease. Ageing Res Rev. 2002, 1 (1): 95-111. 10.1016/S0047-6374(01)00365-7.

Weston PM, King RF, Goode AW, Williams NS: Diet-induced thermogenesis in patients with gastrointestinal cancer cachexia. Clin Sci (Lond). 1989, 77 (2): 133-8.

Webster MJ: Physiological and performance responses to supplementation with thiamin and pantothenic acid derivatives. Eur J Appl Physiol Occup Physiol. 1998, 77 (6): 486-91. 10.1007/s004210050364.

Beek van der EJ, Lowik MR, Hulshof KF, Kistemaker C: Combinations of low thiamin, riboflavin, vitamin B6 and vitamin C intake among Dutch adults. (Dutch Nutrition Surveillance System). J Am Coll Nutr. 1994, 13 (4): 383-91.

Beek van der EJ: Vitamin supplementation and physical exercise performance. J Sports Sci. 1991, 77-90. Spec No

Pedersen BK, Bruunsgaard H, Jensen M, Krzywkowski K, Ostrowski K: Exercise and immune function: effect of ageing and nutrition. Proc Nutr Soc. 1999, 58 (3): 733-42.

Petersen EW, Ostrowski K, Ibfelt T, Richelle M, Offord E, Halkjaer-Kristensen J, Pedersen BK: Effect of vitamin supplementation on cytokine response and on muscle damage after strenuous exercise. Am J Physiol Cell Physiol. 2001, 280 (6): C1570-5.

Grados F, Brazier M, Kamel S, Duver S, Heurtebize N, Maamer M, Mathieu M, Garabedian M, Sebert JL, Fardellone P: Effects on bone mineral density of calcium and vitamin D supplementation in elderly women with vitamin D deficiency. Joint Bone Spine. 2003, 70 (3): 203-8. 10.1016/S1297-319X(03)00046-0.

Brutsaert TD, Hernandez-Cordero S, Rivera J, Viola T, Hughes G, Haas JD: Iron supplementation improves progressive fatigue resistance during dynamic knee extensor exercise in iron-depleted, nonanemic women. Am J Clin Nutr. 2003, 77 (2): 441-8.

Bohl CH, Volpe SL: Magnesium and exercise. Crit Rev Food Sci Nutr. 2002, 42 (6): 533-63. 10.1080/20024091054247.

Lukaski HC: Magnesium, zinc, and chromium nutrition and athletic performance. Can J Appl Physiol. 2001, 26 (Suppl): S13-22.

Morton DP, Callister R: Characteristics and etiology of exercise-related transient abdominal pain. Med Sci Sports Exerc. 2000, 32 (2): 432-8. 10.1097/00005768-200002000-00026.

Noakes TD: Fluid and electrolyte disturbances in heat illness. Int J Sports Med. 1998, 19 (Suppl 2): S146-9. 10.1055/s-2007-971982.

Margaritis I, Tessier F, Prou E, Marconnet P, Marini JF: Effects of endurance training on skeletal muscle oxidative capacities with and without selenium supplementation. J Trace Elem Med Biol. 1997, 11 (1): 37-43.

Tessier F, Margaritis I, Richard MJ, Moynot C, Marconnet P: Selenium and training effects on the glutathione system and aerobic performance. Med Sci Sports Exerc. 1995, 27 (3): 390-6.

McCutcheon LJ, Geor RJ: Sweating. Fluid and ion losses and replacement. Vet Clin North Am Equine Pract. 1998, 14 (1): 75-95.

Gibson RS, Heath AL, Ferguson EL: Risk of suboptimal iron and zinc nutriture among adolescent girls in Australia and New Zealand: causes, consequences, and solutions. Asia Pac J Clin Nutr. 2002, 11 (Suppl 3): S543-52. 10.1046/j.1440-6047.11.supp3.10.x.

Singh A, Failla ML, Deuster PA: Exercise-induced changes in immune function: effects of zinc supplementation. J Appl Physiol. 1994, 76 (6): 2298-303.

Download references

Acknowledgements

The authors would like to thank all of the research participants, graduate students, and researchers that contributed to the body of research cited in this comprehensive review. The authors would like to thank Mr. Chris Noonan for reviewing definition and regulation of dietary supplement section. This article was reviewed and approved by the Research Committee of the ISSN and therefore can be viewed as the official position of the ISSN. Individuals interested in trying some of these nutritional recommendations should do so only after consulting with their personal physician.

Author information

Authors and affiliations.

Exercise & Sports Nutrition Lab, Texas A&M University, College Station, TX, USA

Richard B Kreider

Exercise & Sport Science Department, University of Mary-Hardin Baylor, Belton, TX, USA

Colin D Wilborn & Lem Taylor

School of Physical Education & Exercise Science, University of South Florida, Tampa, FL, USA

Bill Campbell

GENr8, Inc, Dana Point, CA, USA

Anthony L Almada

Collins, McDonald & Gann, PC, Mineola, NY, USA

Rick Collins

Schools of Medicine & Health Movement Studies, The University of Queensland, Herston, Queensland, AU

Mathew Cooke

Pennington Biomedical Reseach Center, Baton Rouge, LA, USA

Conrad P Earnest

Department of Health, Human Performance, and Recreation, Baylor University, Box 97313, Waco, TX, USA

Mike Greenwood, Brian Leutholtz & Darryn S Willoughby

Miami Research Associates, Miami, FL, USA

Douglas S Kalman

Department of Health and Exercise Science, University of Oklahoma, Norman, OK, USA

Chad M Kerksick & Abbie Smith

High Performance Nutrition LLC, Mercer Island, WA, USA

Susan M Kleiner

Northwestern University Feinberg School of Medicine, Department of Physical Medicine and Rehabilitation, Rehabilitation Institute of Chicago, Chicago, IL, USA

Hector Lopez

Nutrition Assessment Laboratory, Nutrition Center, University of Akron, Akron, Ohio, USA

Lonnie M Lowery

Department of Human Performance & Sport Management, Mount Union College, Alliance, Ohio, USA

Marie Spano Nutrition Consulting, Atlanta, GA, USA

Marie Spano

Department of Human Nutrition, Kansas State University, Manhattan, KS, USA

Robert Wildman

Division of Sports Nutrition and Exercise Science, the Center for Applied Health Sciences, Fairlawn, OH, USA

Tim N Ziegenfuss

Farquhar College of Arts and Sciences, Nova Southeastern University, Fort Lauderdale, FL, USA

Jose Antonio

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Richard B Kreider .

Additional information

Competing interests.

Authors of this paper have not received any financial remuneration for preparing or reviewing this paper. However, in an interest of full-disclosure as recommended in this paper, authors report the following competing interests. RBK has received university-funded grants to conduct research on several nutrients discussed in this paper and currently receives research funding from Curves International, General Mills Bell Institute for Human Nutrition; and, the National Institutes of Health. In addition, he has served as a paid consultant for industry; is currently serving as a product development consultant for Supreme Protein, has received honoraria for speaking at conferences and writing lay articles about topics discussed in this paper; receives royalties from the sale of several exercise and nutrition-related books; and, has served as an expert witness on behalf of the plaintiff and defense in cases involving dietary supplements. CW has received academic and industry funding related to dietary supplements and honoraria from speaking engagements on the topic. LT has received academic and industry funding related to dietary supplements and honoraria from speaking engagements on the topic. BC has received university and private sector funded grants to conduct research on several nutrients discussed in this paper and has received compensation for speaking at conferences and writing lay articles/books about topics discussed in this paper. ALA has received consulting fees from AquaGenus, Bergstrom Nutrition, Bioiberica, Curves International, Indena, Indfrag, Miami Research Associates, Omniactives, Sabinsa, and Yor Health; received dietary ingredient materials from Alzchem, Glanbia, and Lonza; sits on the board of New Era; has executive positions in Fein Innovations, Fierce Foods, and GENr8; has equity in AquaGenus, Fein Innovations, Fierce Foods, and GENr8; has stock options in New Era Nutrition and Scientific Food Solutions; has received royalties from Isatori; is a lead inventor on a patent pending related to vitamin K and MSM; has received travel and lodging reimbursement from Bergstrom, Danisco, Indfrag, and New Era Nutrition; has received in-kind compensation from Advanced Research Press; and is on the editorial advisory board of Nutrition Business Journal , and is a columnist for Nutraceuticals World and Muscular Development . RC is the attorney for numerous companies in the dietary supplement industry and has received payment for consultancy and the writing of lay articles discussing nutritional supplements. MC has served as a consultant for industry and received honoraria for speaking about topics discussed in this paper. CPE received honoraria from scientific and lay audience speaking engagements; has served as an expert witness for several patent litigations involving dietary supplements on the behalf of the plaintiff and defense; and, currently has a grant from the Gatorade Sports Science Institute involving the examination of a dietary supplement and its effect on athletic performance. MG has received academic and industry funding to conduct sport/exercise nutritional supplement research; has served as a paid consultant for the sports nutrition industry; and, has received honoraria for speaking engagements and publishing articles in lay sport nutrition venues. DSK has received grants and contracts to conduct research on several nutrients discussed in this paper; has served as a paid consultant for industry; has received honoraria for speaking at conferences and writing lay articles about topics discussed in this paper; receives royalties from the sale of several exercise and nutrition-related books; and, has served as an expert witness on behalf of the plaintiff and defense in cases involving dietary supplements. CMK has received academic and industry funding related to dietary supplements and honoraria from speaking engagements on the topic. In addition, he has received payment for writing of lay articles discussing nutritional supplements. SMK has served as a paid consultant for industry; has received honoraria for speaking at conferences and writing lay articles about topics discussed in this paper; receives royalties from the sale of several exercise and nutrition related books; and, receives commission and has stock in companies that sell products produced from several ingredients discussed in this paper. HL reports having received honoraria for lectures from scientific, educational and community groups; serving as a consultant and scientific advisory board member for Nordic Naturals, Inc.; payment for scientific and technical writing for Optimal Aging and Aesthetic Medicine, LLC.; payment for commercial writing for Essentials of Healthy Living; consultancy fees as owner of Physicians Pioneering Performance, LLC.; owner and medical director of Performance Spine and Sports Medicine, LLC.; and, owner and medical director of Northeast Spine and Sports Medicine, PC. LML has received academic and industry funding related to dietary supplements and honoraria from speaking engagements on the topic and has received payment for consultancy and the writing of lay articles discussing nutritional supplements. RM has received industry funding and stock options related to dietary supplement research. RM has also received honoraria for speaking and payment for consultancy and the writing of lay articles on nutritional supplements. AS reports no competing interests. MS has received honoraria from academic organizations for speaking at conferences and writing lay articles on various sports nutrition topics. TNZ has received university and contract research organization-funded grants to conduct research on several ingredients discussed in this paper; has served as a paid consultant for the sports nutrition industry; has received honoraria for speaking at conferences and writing lay articles about topics discussed in this paper; has received royalties from the sale of dietary supplements; has stock in a company that sells several ingredients discussed in this paper; and, has served as an expert witness in cases involving dietary supplements. RW has received industry funds for consultancy and employment related to dietary supplement development and marketing. DSW has received university and contract research organization-funded grants to conduct research on several ingredients discussed in this paper. He has previously served as a paid consultant for the nutraceutical and sports nutrition industry with the companies, Amino Vital and Transformation Enzyme, and is presently a paid consultant for VPX. He has received honoraria for speaking at conferences and writing lay articles about topics discussed in this paper. JA is the CEO of the ISSN and has received academic and industry (i.e. VPX/Redline) funding related to dietary supplement consultation, speaking engagements and writing on the topic.

Authors' contributions

RBK contributed most of the content and served as senior editor of the paper. CDW, LT, and BC updated references, updated several sections of the paper, and assisted in editing content. ALA, RC, MC, CPE, MG, DSK, CMK, SMK, BL, HL, LML, RM, AS, MS, RW, DSW, TNZ, and JA reviewed and edited the manuscript. All authors read and approved the final manuscript.

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License ( http://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article.

Kreider, R.B., Wilborn, C.D., Taylor, L. et al. ISSN exercise & sport nutrition review: research & recommendations. J Int Soc Sports Nutr 7 , 7 (2010). https://doi.org/10.1186/1550-2783-7-7

Download citation

Received : 18 January 2010

Accepted : 02 February 2010

Published : 02 February 2010

DOI : https://doi.org/10.1186/1550-2783-7-7

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Sports Nutrition
  • Ergogenic Value
  • Dietary Supplement Health And Education Act (DSHEA)
  • Creatine Monohydrate

Journal of the International Society of Sports Nutrition

ISSN: 1550-2783

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

sports nutrition research paper

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • J Int Soc Sports Nutr

ISSN exercise & sports nutrition review update: research & recommendations

Chad m. kerksick.

1 Exercise and Performance Nutrition Laboratory, School of Health Sciences, Lindenwood University, St. Charles, MO USA

Colin D. Wilborn

2 Exercise & Sport Science Department, University of Mary-Hardin Baylor, Belton, TX USA

Michael D. Roberts

3 School of Kinesiology, Auburn University, Auburn, AL USA

Abbie Smith-Ryan

4 Department of Exercise and Sport Science, University of North Carolina, Chapel Hill, NC USA

Susan M. Kleiner

5 High Performance Nutrition LLC, Mercer Island, WA USA

Ralf Jäger

6 Increnovo, LLC, Milwaukee, WI USA

Rick Collins

7 Collins Gann McCloskey and Barry PLLC, Mineola, NY USA

Mathew Cooke

8 Department of Health and Medical Sciences, Swinburne University of Technology, Hawthorn, Victoria Australia

Jaci N. Davis

Elfego galvan.

9 University of Texas Medical Branch, Galveston, TX USA

Mike Greenwood

10 Exercise & Sports Nutrition Lab, Human Clinical Research Facility, Texas A&M University, College Station, TX USA

Lonnie M. Lowery

11 Department of Human Performance & Sport Business, University of Mount Union, Alliance, OH USA

Robert Wildman

12 Dymatize Nutrition, LLC, Dallas, TX USA

Jose Antonio

13 Department of Health and Human Performance, Nova Southeastern University, Davie, FL USA

Richard B. Kreider

Sports nutrition is a constantly evolving field with hundreds of research papers published annually. In the year 2017 alone, 2082 articles were published under the key words ‘sport nutrition’. Consequently, staying current with the relevant literature is often difficult.

This paper is an ongoing update of the sports nutrition review article originally published as the lead paper to launch the Journal of the International Society of Sports Nutrition in 2004 and updated in 2010. It presents a well-referenced overview of the current state of the science related to optimization of training and performance enhancement through exercise training and nutrition. Notably, due to the accelerated pace and size at which the literature base in this research area grows, the topics discussed will focus on muscle hypertrophy and performance enhancement. As such, this paper provides an overview of: 1.) How ergogenic aids and dietary supplements are defined in terms of governmental regulation and oversight; 2.) How dietary supplements are legally regulated in the United States; 3.) How to evaluate the scientific merit of nutritional supplements; 4.) General nutritional strategies to optimize performance and enhance recovery; and, 5.) An overview of our current understanding of nutritional approaches to augment skeletal muscle hypertrophy and the potential ergogenic value of various dietary and supplemental approaches.

Conclusions

This updated review is to provide ISSN members and individuals interested in sports nutrition with information that can be implemented in educational, research or practical settings and serve as a foundational basis for determining the efficacy and safety of many common sport nutrition products and their ingredients.

Evaluating the scientific merit of articles and advertisements about exercise and nutrition products is a key skill that all sports nutrition professionals must possess. To assist members and other advocates of the International Society of Sports Nutrition (ISSN) in keeping up to date about the latest findings in sports nutrition, the ISSN Exercise & Sports Nutrition Review: Research & Recommendations has been updated. The initial version of this paper was the first publication used to help launch the Journal of the International Society of Sports Nutrition (JISSN, originally called the Sports Nutrition Review Journal). This paper provides a definition of ergogenic aids and dietary supplements and discusses how dietary supplements are legally regulated. Other sections highlight how to evaluate the scientific merit of nutritional supplements and provide general nutritional strategies to optimize performance and enhance recovery. Finally, a brief overview of the efficacy surrounding many supplements commonly touted to promote skeletal muscle hypertrophy and improve physical performance is provided. Based upon the available scientific literature testing the efficacy and safety of the nutritional supplements discussed herein, all nutritional supplements discussed in this paper have been placed into three categories based upon the quality and quantity of scientific support available:

  • A) Strong Evidence to Support Efficacy and Apparently Safe
  • B) Limited or Mixed Evidence to Support Efficacy
  • C) Little to No Evidence to Support Efficacy and/or Safety

Since the last published version of this document in 2010 [ 1 ], the general approach to categorization has not changed, but several new supplements have been introduced to the market and are subsequently reviewed in this article. In this respect, many supplements have had additional studies published that has led to some supplements being placed into a different category or removed from the review altogether. We understand and expect that some individuals may not agree with our interpretations of the literature or what category we have assigned a particular supplement, but it is important to appreciate that some classifications may change over time as more research becomes available.

Definition of an ergogenic aid

An ergogenic aid is any training technique, mechanical device, nutritional ingredient or practice, pharmacological method, or psychological technique that can improve exercise performance capacity or enhance training adaptations [ 2 – 4 ]. Ergogenic aids may help prepare an individual to exercise, improve exercise efficiency, enhance recovery from exercise, or assist in injury prevention during intense training. Although this definition seems rather straightforward, there is considerable debate regarding the ergogenic value of various nutritional supplements. A consensus exists to suggest that a nutritional supplement is ergogenic if peer-reviewed studies demonstrate the supplement significantly enhances exercise performance following weeks to months of ingestion (e.g., promotes increases in maximal strength, running speed, and/or work during a given exercise task). On the other hand, a supplement may also have ergogenic value if it acutely enhances the ability of an athlete to perform an exercise task or enhances recovery from a single exercise bout. The ISSN has adopted a broader view regarding the ergogenic value of supplements. While the muscle building and performance enhancing effects of a supplement on a single bout of exercise may lead to eventual ergogenic effects or optimized training adaptations, our view is that such evidence does not warrant “Excellent Evidence to Support Efficacy” if there is a lack of long-term efficacy data. Herein, we have adopted the view that a supplement is clearly ergogenic if most of human studies support the ingredient as being effective in promoting further increases in muscle hypertrophy or performance with exercise training. Conversely, supplements that fall short of this standard and are only supported by preclinical data (e.g., cell culture or rodent studies) are grouped into other categories.

Definition and regulation of dietary supplements

The dietary supplement health and education act (dshea) and the safety of dietary supplements.

Congress passed the Dietary Supplement Health and Education Act of 1994 (DSHEA), placing dietary supplements in a special category of “foods”. In October 1994, President Clinton signed DSHEA into law. This statute was enacted amid claims that the Food and Drug Administration (FDA) was distorting the then-existing provisions of the Food, Drug, and Cosmetic Act (FDCA) to improperly deprive the public of safe and popular dietary supplement products.

The law defines a “dietary supplement” as a product that is intended to supplement the diet and contains a “dietary ingredient”. By definition, “dietary ingredients” in these products may include vitamins, minerals, herbs or other botanicals, amino acids, and substances such as enzymes, organ tissues, and glandular extracts. Further, dietary ingredients may also include extracts, metabolites, or concentrates of those substances. Dietary supplements may be found in many forms such as tablets, capsules, softgels, gelcaps, liquids, or powders, but may only be intended for oral ingestion. Dietary supplements cannot be marketed or promoted for sublingual, intranasal, transdermal, injected, or in any other route of administration except oral ingestion. A supplement can be in other forms, such as a bar, as long as the information on its label does not represent the product as a conventional food or a sole item of a meal or diet.

Indeed, DSHEA clearly defines “dietary supplements” and “dietary ingredients,” it sets certain criteria for “new dietary ingredients,” and the law prevents the FDA from overreaching. Additionally, and contrary to widespread opinion, DSHEA did not leave the industry unregulated. The dietary supplement industry is in fact regulated by the FDA as a result of DSHEA. The law ensures the authority of the FDA to provide legitimate protections for the public health. The Federal Trade Commission (FTC) also continues to have jurisdiction over the marketing claims that dietary supplement manufacturers or companies make about their products. The FDA and FTC operate in a cooperative fashion to regulate the dietary supplement industry. In this respect, the extent to which information is shared and jurisdiction between these two entities overlaps with regard to marketing and advertising dietary supplements continues to increase.

In the United States, dietary supplements are classified as food products, not drugs, and there is generally no mandate to register products with the FDA or obtain FDA approval before producing or selling supplements to consumers. However, if a dietary supplement manufacturer is making a claim about their product, the company must submit the claims to FDA within 30 days of marketing the product. Compare this, for example, with Canada where under the Natural Health Product (NHP) Regulations enacted in 2004 supplements must be reviewed, approved, and registered with Health Canada. The rationale for the U.S. model is based on a presumed long history of safe use; hence there is no need to require additional safety data.

DSHEA also requires supplement marketers to include on any label displaying structure/function claims (i.e., claims that the product affects the structure or function of the body) the mandatory FDA disclaimer “This statement has not been evaluated by the Food and Drug Administration. This product is not intended to diagnose, treat, cure, or prevent any disease.” Opponents of dietary supplements often cite this statement as evidence that the FDA does not review or approve dietary supplements. However, most dietary ingredients have been “grandfathered in” as DSHEA-compliant ingredients due to a long history of safe use, and those products containing new ingredients must be submitted by a notification to the FDA for a safety review prior to being brought to market. Although many dietary ingredients have been introduced into dietary supplements since October 1994 and have not been submitted to the FDA for a safety review, nutritional supplementation writ large is generally safe. In this regard, while there are over 50,000 dietary supplements registered with the Office of Dietary Supplement’s “Dietary Supplement Label Database”, a 2013 Annual Report (released in 2015) of the American Association of Poison Control Centers revealed zero fatalities occurred due to dietary supplements compared to 1692 deaths due to drugs [ 5 ]. Perhaps more alarming is a 2015 report by the Centers for Disease Control suggesting 2,287,273 emergency room visits were due to prescription drug-related events which dwarfs the 3266 emergency room visits due to dietary supplements (adjusted from 23,000 visits after excluding cases of older adults choking on pills, allergic reactions, unsupervised children consuming too many vitamins, and persons consuming ingredients not defined by DSHEA as a dietary supplement) [ 5 ]. Furthermore, a recent Healthcare Cost and Utilization Project Statistical Brief by Lucado et al. [ 6 ] reported approximately one in six Americans suffered from food borne illnesses in 2010, and food borne illnesses were associated with over 3.7 million treat-and-release emergency department visits, 1.3 million inpatient hospital stays, and 3000 deaths. Notwithstanding, there have been case reports of liver and kidney toxicity potentially caused by supplements containing herbal extracts [ 7 ] as well as overdoses associated with pure caffeine anhydrous ingestion [ 8 ]. Collectively, the aforementioned statistics and case reports demonstrate that while generally safe, as with food or prescription drug consumption, dietary supplement consumption can lead to adverse events in spite of DSHEA and current FDA regulations described below.

New dietary ingredients

Recognizing that new and untested dietary supplement products may pose unknown health issues, DSHEA distinguishes between products containing dietary ingredients that were already on the market and products containing new dietary ingredients that were not marketed prior to the enactment of the law. A “new dietary ingredient” (NDI) is defined as a dietary ingredient that was not marketed in the United States before October 15, 1994. DSHEA grants the FDA greater control over supplements containing NDIs. A product containing an NDI is deemed adulterated and subject to FDA enforcement sanctions unless it meets one of two exemption criteria: either (1) the supplement in question contains “only dietary ingredients which have been present in the food supply as an article used for food in a form in which the food has not been chemically altered”; or (2) there is a “history of use or other evidence of safety” provided by the manufacturer or distributor to the FDA at least 75 days before introducing the product into interstate commerce. The first criterion is silent as to how and by whom presence in the food supply as food articles without chemical alteration is to be established. The second criterion—applicable only to new dietary ingredients that have not been present in the food supply—requires manufacturers and distributors of the product to take certain actions. Those actions include submitting, at least 75 days before the product is introduced into interstate commerce, information that is the basis on which a product containing the new dietary ingredient is “reasonably expected to be safe.” That information would include: (1) the name of the new dietary ingredient and, if it is an herb or botanical, the Latin binomial name; (2) a description of the dietary supplement that contains the new dietary ingredient, including (a) the level of the new dietary ingredient in the product, (b) conditions of use of the product stated in the labeling, or if no conditions of use are stated, the ordinary conditions of use, and (c) a history of use or other evidence of safety establishing that the dietary ingredient, when used under the conditions recommended or suggested in the labeling of the dietary supplement, is reasonably expected to be safe.

In July 2011, the FDA released a Draft Guidance for Industry, entitled “Dietary Supplements: New Dietary Ingredient Notifications and Related Issues.” While a guidance does not carry the authority or the enforceability of a law or regulation, the FDA’s NDI draft guidance represented the agency’s current thinking on the topic. The guidance prompted great controversy, and FDA agreed to issue a revised draft guidance to address some of the issues raised by industry. In August 2016, FDA released a revised Draft Guidance that replaced the 2011 Draft Guidance. The purpose of the 2016 Draft Guidance was to help manufacturers and distributors decide whether to submit a premarket safety notification to FDA, help prepare NDI notifications in a manner that allows FDA to review and respond more efficiently and quickly, and to improve the quality of NDI notifications. The 2016 Draft Guidance has been criticized by industry and trade associations for its lack of clarity and other problems. Some of these issues include the lack of clarity regarding Pre-DSHEA, (Grandfathered), ingredients and FDA requiring an NDI notification even if another manufacturer has submitted a notification.

The lack of clarity surrounding the “new” Draft Guidance has led to many NDI notifications being rejected by FDA for lack of safety data and other issues. Other companies have opted to utilize the “Self-Affirmed Generally Recognized as Safe (GRAS)” route in order to “bypass” the NDI notification process. Self-Affirmed GRAS is when a company has a team of scientific experts evaluate the safety of their ingredient. There is no requirement that the safety dossier be submitted to FDA but is used by the company as an internal document that may be relied upon if the ingredient is challenged by the FDA. FDA has expressed its concern with this practice and does not encourage dietary supplement manufacturers to use Self-Affirmed GRAS to avoid submitting NDI notifications. In any event, the likelihood of another revised Draft Guidance from FDA becoming available in the future is high, and possibly more enforcement actions taken against companies that market an NDI without submitting a notification.

Adverse event reporting

In response to growing criticism of the dietary supplement industry, the 109th Congress passed the first mandatory Adverse Event Reporting (AER) legislation for the dietary supplement industry. In December 2006, President Bush signed into law the Dietary Supplement and Nonprescription Drug Consumer Protection Act, which took effect on December 22, 2007. After much debate in Congress and input from the FDA, the American Medical Association (AMA), many of the major supplement trade associations, and a host of others all agreed that the legislation was necessary and the final version was approved by all. In short, the Act requires that all “serious adverse events” regarding dietary supplements be reported to the Secretary of Health and Human Services. The law strengthens the regulatory structure for dietary supplements and builds greater consumer confidence, as consumers have a right to expect that if they report a serious adverse event to a dietary supplement marketer the FDA will be advised about it.

An adverse event is any health-related event associated with the use of a dietary supplement that is adverse. A serious adverse event is an adverse event that (A) results in (i) death, (ii) a life-threatening experience, (iii) inpatient hospitalization, (iv) a persistent or significant disability or incapacity, or (v) a congenital anomaly or birth defect; or (B) requires, based on reasonable medical judgment, a medical or surgical intervention to prevent an outcome described under subparagraph (A). Once it is determined that a serious adverse event has occurred, the manufacturer, packer, or distributor (responsible person) of a dietary supplement whose name appears on the label of the supplement shall submit to the Secretary of Health and Human Services any report received of the serious adverse event accompanied by a copy of the label on or within the retail packaging of the dietary supplement. The responsible person has 15 business days to submit the report to FDA after being notified of the serious adverse event. Following the initial report, the responsible person must submit follow-up reports of new medical information that they receive for one-year.

Adulterated supplements

The FDA has various options to protect consumers from unsafe supplements. The Secretary of the Department of Health and Human Services (which falls under FDA oversight) has the power to declare a dangerous supplement to be an “imminent hazard” to public health or safety and immediately suspend sales of the product. The FDA also has the authority to protect consumers from dietary supplements that do not present an imminent hazard to the public but do present certain risks of illness or injury to consumers. The law prohibits introducing adulterated products into interstate commerce. A supplement shall be deemed adulterated if it presents “a significant or unreasonable risk of illness or injury”. The standard does not require proof that consumers have actually been harmed or even that a product will harm anyone. It was under this provision that the FDA concluded that dietary supplements containing ephedra, androstenedione, and DMAA presented an unreasonable risk. Most recently, FDA imposed an importation ban on the botanical Mitragyna speciose, better known as Kratom. In 2016, FDA issued Import Alert #54–15, which allows for detention without physical examination of dietary supplements and bulk dietary ingredients that are, or contain, Kratom. Criminal penalties are present for a conviction of introducing adulterated supplement products into interstate commerce. While the harms associated with dietary supplements may pale in comparison to those linked to prescription drugs, recent pronouncements from the U.S. Department of Justice confirm that the supplement industry is being watched vigilantly to protect the health and safety of the American public.

Good manufacturing practices

When DSHEA was passed in 1994, it contained a provision requiring that the FDA establish and enforce current Good Manufacturing Practices (cGMPs) for dietary supplements. However, it was not until 2007 that the cGMPs were finally approved, and not until 2010 that the cGMPs applied across the industry, to large and small companies alike. The adherence to cGMPs has helped protect against contamination issues and should serve to improve consumer confidence in dietary supplements. The market improved as companies became compliant with cGMPs, as these regulations imposed more stringent requirements such as Vendor Certification, Document Control Procedures, and Identity Testing. These compliance criteria addressed the problems that had damaged the reputation of the industry with a focus on quality control, record keeping, and documentation.

However, it does appear that some within the industry continue to struggle with compliance. In Fiscal Year 2017, it was reported that approximately 23.48% of the FDA’s 656 total cGMP inspections resulted in citations for failing to establish specifications for the identity, purity, strength, and composition of dietary supplements. Further, 18.47% of those inspected were cited for failing to establish and/or follow written procedures for quality control operations. Undoubtedly, relying on certificates of analysis from the raw materials supplier without further testing, or failing to conduct identity testing of a finished product, can result in the creation of a product that contains something it should not contain such as synthetic chemicals or even pharmaceutical drugs. All members of the industry need to ensure compliance with cGMPs.

Marketing claims

According to the 1990 Nutrition Labeling and Education Act (NLEA), the FDA can review and approve health claims (claims describing the relationship between a food substance and a reduced risk of a disease or health-related condition) for dietary ingredients and foods. However, since the law was passed it has only approved a few claims. The delay in reviewing health claims of dietary supplement ingredients resulted in a lawsuit, Pearson v. Shalala , filed in 1995. After years of litigation, in 1999 the U.S. Court of Appeals for the District of Columbia Circuit ruled that qualified health claims may be made about dietary supplements with approval by FDA, as long as the statements are truthful and based on adequate science. Supplement or food companies wishing to make health claims or qualified health claims about supplements can submit research evidence to the FDA for review.

The FTC also regulates the supplement industry. Unsubstantiated claims invite enforcement by the FTC (along with the FDA, state district attorney offices, groups like the Better Business Bureau, and plaintiff’s lawyers who file class action lawsuits). The FTC has typically applied a substantiation standard of “competent and reliable scientific evidence” to claims about the benefits and safety of dietary supplements. FTC case law defines “competent and reliable scientific evidence” as “tests, analyses, research, studies, or other evidence based on the expertise of professionals in the relevant area, that has been conducted and evaluated in an objective manner by persons qualified to do so, using procedures generally accepted in the profession to yield accurate and reliable results.” The FTC has claimed that this involves providing at least two clinical trials showing efficacy of the actual product, within a population of subjects relevant to the target market, supporting the structure/function claims that are made. While the exact requirements are still evolving, the FTC has acted against several supplement companies for misleading advertisements and/or structure/function claims.

A safer industry ahead

As demonstrated, while some argue that the dietary supplement industry is “unregulated” and/or may have suggestions for additional regulation, manufacturers and distributors of dietary supplements must adhere to several federal regulations before a product can go to market. Further, before marketing products, they must have evidence that their supplements are generally safe to meet all the requirements of DSHEA and FDA regulations. For this reason, over the last 20 years, many established supplement companies have employed research and development directors who help educate the public about nutrition and exercise, provide input on product development, conduct preliminary research on products, and/or assist in coordinating research trials conducted by independent research teams (e.g., university-based researchers or clinical research sites). These companies also consult with marketing and legal teams with the responsibility to ensure structure/function claims do not misrepresent results of research findings. This has increased job opportunities for sports nutrition specialists as well as enhanced external funding opportunities for research groups interested in exercise and nutrition research.

While some companies have falsely attributed research on different dietary ingredients or dietary supplements to their own products, suppressed negative research findings, and/or exaggerated results from research studies, the trend in the sports nutrition industry has been to develop scientifically sound supplements. This trend toward greater research support is the result of: (1) attempts to honestly and accurately inform the public about results; (2) efforts to obtain data to support safety and efficacy on products for the FDA and the FTC; and/or, (3) endeavors to provide scientific evidence to support advertising claims and increase sales. While the push for more research is due in part to greater scrutiny from the FDA and FTC, it is also in response to an increasingly competitive marketplace where established safety and efficacy attracts more consumer loyalty and helps ensure a longer lifespan for the product in commerce. Companies that adhere to these ethical standards tend to prosper while those that do not will typically struggle to comply with FDA and FTC guidelines resulting in a loss of consumer confidence and an early demise for the product.

Product development and quality assurance

A common question posed by athletes, parents, and professionals surrounding dietary supplements relates to how they are manufactured and perceived supplement quality. In several cases, established companies who develop dietary supplements have research teams who scour the medical and scientific literature looking for potentially effective nutrients. These research teams often attend scientific meetings and review the latest patents, research abstracts presented at scientific meetings, and research publications. Leading companies invest in basic research on nutrients before developing their supplement formulations and often consult with leading researchers to discuss ideas about dietary supplements and their potential for commercialization. Other companies wait until research has been presented in patents, research abstracts, or publications before developing nutritional formulations featuring the nutrient. Upon identification of new nutrients or potential formulations, the next step is to contact raw ingredient suppliers to see if the nutrient is available, if it is affordable, how much of it can be sourced and what is the available purity. Sometimes, companies develop and pursue patents involving new processing and purification processes because the nutrient has not yet been extracted in a pure form or is not available in large quantities. Reputable raw material manufacturers conduct extensive tests to examine purity of their raw ingredients. When working on a new ingredient, companies often conduct series of toxicity studies on the new nutrient once a purified source has been identified. The company would then compile a safety dossier and communicate it to the FDA as a New Dietary Ingredient submission, with the hopes of it being allowed for lawful sale.

When a powdered formulation is designed, the list of ingredients and raw materials are typically sent to a flavoring house and packaging company to identify the best way to flavor and package the supplement. In the nutrition industry, several main flavoring houses and packaging companies exist who make many dietary supplements for supplement companies. Most reputable dietary supplement manufacturers submit their production facilities to inspection from the FDA and adhere to GMP, which represent industry standards for good manufacturing of dietary supplements. Some companies also submit their products for independent testing by third-party companies to certify that their products meet label claims and that the product is free of various banned ingredients. For example, the certification service offered by NSF International includes product testing, GMP inspections, ongoing monitoring and use of the NSF Mark indicating products comply with inspection standards, and screening for contaminants. More recently, companies have subjected their products for testing by third party companies to inspect for banned or unwanted substances. These types of tests help ensure that the dietary supplement made available to athletes do not contained substances banned by the International Olympic Committee or other athletic governing bodies (e.g., NFL, NCAA, MLB, NHL, etc.). While third-party testing does not guarantee that a supplement is void of banned substances, the likelihood is reduced (e.g., Banned Substances Control Group, Informed Choice, NSF, etc.). Moreover, consumers can request copies of results of these tests and each product that has gone through testing and earned certification can be researched online to help athletes, coaches and support staff understand which products should be considered. In many situations, companies who are not willing to provide copies of test results or certificates of analysis should be viewed with caution, particularly for individuals whose eligibility to participate might be compromised if a tainted product is consumed.

Evaluation of nutrition ergogenic aids

The ISSN recommends that potential consumers undertake a systematic process of evaluating the validity and scientific merit of claims made when assessing the ergogenic value of a dietary supplement [ 1 , 4 ]. This can be accomplished by examining the theoretical rationale behind the supplement and determining whether there is any well-controlled data showing the supplement is effective. Supplements based on sound scientific rationale with direct, supportive research showing effectiveness may be worth trying or recommending. However, those based on unsound scientific results or offer little to no data supporting the ergogenic value of the actual supplement/technique may not be worthwhile. Sports nutrition specialists should be a resource to help their clients interpret the scientific and medical research that may impact their welfare and help them train more effectively. The following are recommended questions to ask when evaluating the potential ergogenic value of a supplement.

Does the theory make sense?

Most supplements that have been marketed to improve health or exercise performance are based on theoretical applications derived from basic science or clinical research studies. Based on these preliminary studies, a dietary approach or supplement is often marketed to people proclaiming the benefits observed in these basic research studies. Although the theory may appear relevant, critical analysis of this process often reveals flaws in the scientific logic or that the claims made do not quite match up with the cited literature. By evaluating the literature one can discern whether or not a dietary approach or supplement has been based on sound scientific evidence. To do so, one is recommended to first read reviews about the training method, nutrient, or supplement from researchers who have been intimately involved in the available research and consult reliable references about nutritional and herbal supplements [ 1 , 9 ]. To aid in this endeavour, the ISSN has published position statement on topics related to creatine [ 10 ], protein [ 11 ], beta-alanine [ 12 ], nutrient timing [ 13 ], caffeine [ 14 ], HMB [ 15 ], meal frequency [ 16 ], energy drinks [ 17 ], and diets and body composition [ 18 ]. Each of these documents would be excellent resources for any of these topics. In addition, other review articles and consensus statements have been published by other researchers and research groups that evaluate dietary supplements, offer recommendations on interpreting the literature, and discuss the available findings for several ingredients that are discussed in this document [ 19 – 21 ]. We also advise consumers to conduct a search on the nutrient, key ingredients or the supplement itself on the National Library of Medicine’s Pub Med Online ( https://www.ncbi.nlm.nih.gov/pubmed/ ). A quick look at these references will often help determine if the theoretical impetus for supplementing with an ingredient is plausible or not. Proponents of ergogenic aids often overstate claims made about training devices and dietary supplements while opponents of ergogenic aids and dietary supplements are often either unaware or are ignorant of research supporting their use. Sports nutrition specialists have the responsibility to know the literature and search available databases to evaluate the level of merit surrounding a proposed ergogenic aid.

Is the supplement legal and safe?

An initial question that should be asked is whether the supplement is legal and/or safe. Some athletic associations have banned the use of various nutritional supplements (e.g., prohormones, ephedra that contains ephedrine, “muscle building” supplements, etc.) and many professional sports organization have now written language into their collective bargaining agreements that products made available by the team must be NSF certified as safe for sport. Obviously, if the supplement is banned, the sports nutrition specialist should discourage its use. In addition, many supplements lack appropriate long-term safety data. People who consider taking nutritional supplements should be well aware of the potential side effects so they can make an informed decision whether to use a supplement. Additionally, they should consult with a knowledgeable physician to see if any underlying medical problems exist that may contraindicate its use. When evaluating the safety of a supplement, it is suggested to determine if any side effects have been reported in the scientific or medical literature. In particular, we suggest determining how long a particular supplement has been studied, the dosages evaluated, and whether any side effects were observed. We also recommend consulting the Physician’s Desk Reference (PDR) for nutritional supplements and herbal supplements to see if any side effects have been reported and/or if there are any known drug interactions. If no side effects have been reported in the scientific/medical literature, we generally will view the supplement as safe for the length of time and dosages evaluated. Unfortunately, many available supplements have not had basic safety studies completed that replicate the length of time and dosages being used.

Is there any scientific evidence supporting the ergogenic value?

The next question to ask is whether any well-controlled data are available showing effectiveness of the proposed ergogenic aid in athletic populations or people regularly involved in exercise training. The first place to look is the list of references cited in marketing material supporting their claims. Are the abstracts or articles cited just general references or specific studies that have evaluated the efficacy of the nutrients included in the formulation or of the actual supplement? From there, one can critically evaluate the cited abstracts and articles by asking a series of questions:

  • Are the studies basic research done in animals/clinical populations or have the studies been conducted on athletes/trained subjects? For perspective, studies reporting improved performance in rats or an individual diagnosed with type 2 diabetes may be insightful, but research conducted on non-diabetic athletes is much more practical and relevant.
  • Were the studies well controlled? For ergogenic aid research, the gold standard study design is a randomized, double-blind, placebo controlled clinical trial. This means that neither the researcher nor the subject is aware which group received the supplement or the placebo during the study and that the subjects were randomly assigned into the placebo or supplement group. An additional element of rigor is called a cross-over design, where each subject, at different times (separated by an interval known as a “washout period”), is exposed to each of the treatments. While utilization of a cross-over design is not always feasible, it reduces the element of variability within a participant and subsequently, increases the strength of study’s findings. At times, supplement claims have been based on poorly designed studies (i.e., small groups of subjects, no control group, use of unreliable tests, etc.) or testimonials which make interpretation more difficult. Well-controlled clinical trials provide stronger evidence as to the potential ergogenic value and importantly how the findings can best be used.
  • Do the studies report statistically significant results or are claims being made on non-significant means or trends? Appropriate statistical analysis of research results allows for an unbiased interpretation of data. Although studies reporting statistical trends may be of interest and lead researchers to conduct additional research, studies reporting statistically significant results are obviously more convincing. With this said, it is important for people to understand that oftentimes the potential effect a dietary supplement or diet regimen may have above and beyond the effect seen from the exercise bout or an accepted dietary approach is quite small. In addition, many studies examining a biochemical or molecular biology mechanism can require invasive sampling techniques or the study population being recruited is unique (very highly trained) resulting in a small number of study participants. When viewed together, the combination of these two considerations can result in statistical outcomes that do not reach statistical significance even though large mean changes were observed. In these situations, the reporting of confidence intervals on the mean change, individual responses from all participants to the investigated treatment and/or effect sizes are additional pieces of information that can allow for a more accurate interpretation. In all such cases, additional research is warranted to further examine the potential ergogenic aid before conclusions can be made.
  • Do the results of the cited studies match the claims made about the supplement or do they accurately portray the response of the supplement against an appropriate placebo or control group? It is not unusual for marketing claims to greatly exaggerate the results found in the actual studies and do so by focusing upon just the outcomes within the supplement (treatment) group as opposed to how the supplement group changed in comparison to how a placebo group changed. Similarly, it is not uncommon for ostensibly compelling results, that may indeed be statistically significant, to be amplified while other relevant findings of significant consumer interest are obscured or omitted (e.g. a dietary supplement showing statistically significant increases in circulating testosterone yet changes in body composition or muscular performance were not superior to a placebo). The only way to determine this is to read the entire article versus focusing an entire study’s interpretation on the provided abstract or even the article citation, and compare results observed in the studies to the available marketing claims. Reputable companies accurately and completely report results of studies so that consumers can make informed decisions about using a product.
  • Were results of the study presented at a reputable scientific meeting and/or published in a peer-reviewed scientific journal? At times, claims are based on research that has either never been published or only published in an obscure journal. The best research is typically presented at respected scientific meetings and/or published in reputable peer-reviewed journals. Three ways to determine a journal’s reputation is either: 1) identify the publisher, 2) the “impact factor” of the journal or 3) whether or not the journal is indexed and subsequently available for review on Pub Med ( https://www.ncbi.nlm.nih.gov/pubmed/ ). Many “peer-reviewed” journals are published by companies with ties to, or are actually owned by, companies that do business with various nutritional products (even though they may be available on PubMed). Therefore, we recommend looking up the publisher’s website and see how many other journals they publish. If you see only a few other journals this is a suggestion that the journal is not a reputable journal. Additionally, one can also look up how many articles have been published by the journal in the last 6–12 months and how many of these articles are well-conducted studies. Alternatively, one can also inquire about the impact factor, a qualitative ranking determined by the number of times a journal’s articles are cited. Impact factors are determined and published by Thomson Reuters under Journal Citation Reports® (a subscription service available at most university libraries). Most journals list their impact factor on the journal home page. Historically, those articles that are read and cited the most are the most impactful scientifically.
  • Have the research findings been replicated? If so, have the results only been replicated at the same laboratory? The best way to know an ergogenic aid works is to see that results have been replicated in several studies preferably by several separate, distinct research groups. The most reliable ergogenic aids are those in which multiple studies, conducted at different labs, have reported similar results of safety and efficacy. Additionally, replication of results by different, unaffiliated labs with completely different authors also removes or reduces the potentially confounding element of publication bias (publication of studies showing only positive results) and conflicts of interest. A notable number of studies on ergogenic aids are conducted in collaboration with one or more research scientists or co-authors that have a real or perceived economic interest in the outcome of the study. This could range from being a co-inventor on a patent application that is the subject of the ergogenic aid, being paid or receiving royalties from the creation of a dietary supplement formulation, providing consulting services for the company or having stock options or shares in a company that owns or markets the ergogenic aid described in the study. An increasing number of journals require disclosures by all authors of scientific articles, and including such disclosures in published articles. This is driven by the aim of providing greater transparency and research integrity. It is important to emphasize that disclosure of a conflict of interest does not alone discredit or dilute the merits of a research study. The primary thrust behind public disclosures of potential conflicts of interest is first and foremost transparency to the reader and second to prevent a later revelation of some form of confounding interest that has the potential of discrediting the study in question, the findings of the study, the authors, and even the research center or institution where the study was conducted.

Classifying and categorizing supplements

Dietary supplements may contain carbohydrate, protein, fat, minerals, vitamins, herbs, enzymes, metabolic intermediates (i.e., select amino acids), or various plant/food extracts. Supplements can generally be classified as convenience supplements (e.g., energy bars, gels, blocks, meal replacement powders, or ready to drink supplements) designed to provide a convenient means of meeting necessary energy or macronutrient needs while also providing support towards attempts at managing caloric intake, weight gain, weight loss, and/or performance enhancement. As discussed previously, evaluating the available scientific literature is an important step in determining the efficacy of any diet, diet program or dietary supplement. In considering this, nutritional supplements can be categorized in the following manner:

  • I. Strong Evidence to Support Efficacy and Apparently Safe: Supplements that have a sound theoretical rationale with the majority of available research in relevant populations using appropriate dosing regimens demonstrating both its efficacy and safety.
  • II. Limited or Mixed Evidence to Support Efficacy: Supplements within this category are characterized as having a sound scientific rationale for its use, but the available research has failed to produce consistent outcomes supporting its efficacy. Routinely, these supplements require more research to be completed before researchers can begin to understand their impact. Importantly, these supplements have no available evidence to suggest they lack safety or should be viewed as harmful.
  • III. Little to No Evidence to Support Efficacy and/or Safety: Supplements within this category generally lack a sound scientific rationale and the available research consistently shows it to lack efficacy. Alternatively, supplements that may be harmful to one’s health or to lack safety are also placed in this category.

Several factors are evaluated when beginning to counsel individuals who regularly complete exercise training. First, a clear understanding of the athlete’s goals and the time with which they have to meet those goals is important. In addition to monitoring load and recovery, an evaluation of the individual’s diet and training program should also be completed. To accomplish this, one should make sure the athlete is eating an energy balanced, nutrient dense diet that meets their estimated daily energy needs and that they are training intelligently. Far too many athletes or coaches focus too heavily upon supplementation or applications of supplementation and neglect these key fundamental aspects. Following this, we suggest that they generally only recommend supplements in category I (i.e., ‘Strong Evidence to Support Efficacy and Apparently Safe’). If an athlete is interested in trying supplements in category II (i.e., ‘Limited or Mixed Evidence to Support Efficacy’), the athlete should make sure they understand these supplements are more experimental and they may or may not see the type of results claimed. Obviously, the ISSN does not support athletes taking supplements in category III (i.e., ‘Little to No Evidence to Support Efficacy and/or Safety’). We believe this approach is scientifically substantiated and offers a balanced view as opposed to simply dismissing the use of all dietary supplements.

General dietary guidelines for active individuals

A well-designed diet that meets energy intake needs and incorporates proper timing of nutrients is the foundation upon which a good training program can be developed [ 22 , 23 ]. Research has clearly shown that lacking sufficient calories and/or enough of the right type of macronutrients may impede an athlete’s training adaptations, while athletes who consume a balanced diet that meets energy needs can augment physiological training adaptations. Moreover, maintaining an energy deficient diet during training may lead to loss of muscle mass, strength, and bone mineral density in addition to an increased susceptibility to illness and injuries, disturbances in immune, endocrine and reproductive function, and an increased prevalence of overreaching and/or overtraining. Incorporating good dietary practices as part of a training program is one way to help optimize training adaptations and prevent overtraining. The following is an overview of energy intake recommendations and major nutrient needs for active individuals.

Energy needs

The primary component to optimize training and performance through nutrition is to ensure the athlete is consuming enough calories to offset energy expenditure [ 22 – 26 ]. People who participate in a general fitness program (e.g., exercising 30–40 min per day, 3 times per week) can typically meet nutritional needs following a normal diet (e.g., 1800–2400 kcals/day or about 25–35 kcals/kg/day for a 50–80 kg individual) because their caloric demands from exercise are not too great (e.g., 200–400 kcals/session). However, athletes involved in moderate levels of intense training (e.g., 2–3 h per day of intense exercise performed 5–6 times per week) or high volume intense training (e.g., 3–6 h per day of intense training in 1–2 workouts for 5–6 days per week) may expend 600–1200 kcals or more per hour during exercise [ 24 ]. For this reason, their caloric needs may approach 40–70 kcals/kg/day (2000–7000 kcals/day for a 50–100 kg athlete). For elite athletes, energy expenditure during heavy training or competition will further exceed these levels [ 27 , 28 ]. For example, energy expenditure for cyclists to compete in the Tour de France has been estimated as high as 12,000 kcals/day (150–200 kcals/kg/day for a 60–80 kg athlete) [ 29 , 30 ]. Additionally, caloric needs for large athletes (i.e., 100–150 kg) may range between 6000 and 12,000 kcals/day depending on the volume and intensity of different training phases [ 31 ].

Although some argue that athletes can meet caloric needs simply by consuming a well-balanced diet, it is often very difficult for larger athletes and athletes engaged in high volume/intense training to be able to eat enough food, on a daily basis, to meet caloric needs [ 2 , 29 , 30 , 32 – 34 ]. This point was clearly highlighted in a review by Burke who demonstrated that carbohydrate needs are largely unmet by high-level athletes [ 22 ]. Additionally it is difficult to consume enough food and maintain gastrointestinal comfort to train or race at peak levels [ 35 ]. Maintaining an energy deficient diet during training often leads to a number of physical (i.e., loss of fat-free mass, illness, reduced sleep quality, incomplete recovery, hormonal fluctuations, increased resting heart rate, etc.) and psychological (i.e., apathy towards training, heightened stress) adverse outcomes [ 23 , 27 ]. Nutritional analyses of athletes’ diets have revealed that many are susceptible to maintaining negative energy intakes during training. It is still a question whether there may be specific individualized occasions when negative energy balance may enhance performance in the days prior to running performance [ 36 ]. Populations susceptible to negative energy balance include runners, cyclists, swimmers, triathletes, gymnasts, skaters, dancers, wrestlers, boxers, and athletes attempting to lose weight too quickly [ 37 ]. Additionally, female athletes are at particular risk of under fueling due to both competitive and aesthetic demands of their sport and their surrounding culture. Female athletes have been reported to have a high incidence of eating disorders [ 38 ]. Low or reduced energy availability (LEA) is linked to functional hypothalamic oligomenorrhea/amenorrhea (FHA), which is frequently reported in weight sensitive sports. This makes LEA a major nutritional concern for female athletes [ 39 ]. Consequently, it is important for the sports nutrition specialist working with athletes to assess athletes individually to ensure that athletes are well fed according to the goals of their sport and their health, and consume enough calories to offset the increased energy demands of training, and maintain body weight. Although this sounds relatively simple, intense training often suppresses appetite and/or alters hunger patterns so that many athletes do not feel like eating [ 37 , 38 ]. Some athletes prefer not to exercise within several hours after eating because of sensations of fullness and/or a predisposition to cause gastrointestinal distress. Further, travel and training schedules may limit food availability or the types of food athletes are accustomed to eating. This means that care should be taken to plan meal times in concert with training, as well as to make sure athletes have sufficient availability of nutrient dense foods throughout the day for snacking between meals (e.g., fluids, carbohydrate/protein-rich foods and supplemental bars, etc.) [ 2 , 33 , 40 ]. For this reason, sports nutritionists’ often recommend that athletes consume four to six meals per day and snacks in between meals to meet energy needs. Due to these practical concerns, the use of nutrient dense energy foods, energy bars and high calorie carbohydrate/protein supplements provides a convenient way for athletes to supplement their diet in order to maintain energy intake during training.

Carbohydrate

Beyond optimal energy intake, consuming adequate amounts of carbohydrate, protein, and fat is important for athletes to optimize their training and performance. In particular and as it relates to exercise performance, the need for optimal carbohydrates before, during and after intense and high-volume bouts of training and competition is evident [ 41 ]. Excellent reviews [ 42 , 43 ] and original investigations [ 44 – 49 ] continue to highlight the known dependence on carbohydrates that exists for athletes competing to win various endurance and team sport activities. A complete discussion of the needs of carbohydrates and strategies to deliver optimal carbohydrate and replenish lost muscle and liver glycogen extend beyond the scope of this paper, but the reader is referred to several informative reviews on the topic [ 23 , 41 , 50 – 53 ].

As such, individuals engaged in a general fitness program and are not necessarily training to meet any type of performance goal can typically meet daily carbohydrate needs by consuming a normal diet (i.e., 45–55% CHO [3–5 g/kg/day], 15–20% PRO [0.8–1.2 g/kg/day], and 25–35% fat [0.5–1.5 g/kg/day]). However, athletes involved in moderate and high-volume training need greater amounts of carbohydrate and protein (discussed later) in their diet to meet macronutrient needs [ 50 ]. In terms of carbohydrate needs, athletes involved in moderate amounts of intense training (e.g., 2–3 h per day of intense exercise performed 5–6 times per week) typically need to consume a diet consisting of 5–8 g/kg/day or 250–1200 g/day for 50–150 kg athletes of carbohydrate to maintain liver and muscle glycogen stores [ 23 , 24 , 50 ]. Research has also shown that athletes involved in high volume intense training (e.g., 3–6 h per day of intense training in 1–2 daily workouts for 5–6 days per week) may need to consume 8–10 g/day of carbohydrate (i.e., 400–1500 g/day for 50–150 kg athletes) in order to maintain muscle glycogen levels [ 50 ]. Preferably, the majority of dietary carbohydrate should come from whole grains, vegetables, fruits, etc. while foods that empty quickly from the stomach such as refined sugars, starches and engineered sports nutrition products should be reserved for situations in which glycogen resynthesis needs to occur at accelerated rates [ 53 ]. In these situations, the absolute delivery of carbohydrate (> 8 g of carbohydrate/kg/day or at least 1.2 g of carbohydrate/kg/hour for the first four hours into recovery) takes precedence over other strategies such as those that may relate to timing or concomitant ingestion of other macronutrients (e.g., protein) or non-nutrients (e.g., caffeine) or carbohydrate type (i.e., glycemic index) [ 50 ].

When considering the carbohydrate needs throughout an exercise session, several key factors should be considered. Previous research has indicated athletes undergoing prolonged bouts (2–3 h) of exercise training can oxidize carbohydrates at a rate of 1–1.1 g per minute or about 60 g per hour [ 41 ]. Several reviews advocate the ingestion of 0.7 g of carbohydrate/kg/hr. during exercise in a 6–8% solution (i.e., 6–8 g per 100 ml of fluid) [ 41 , 42 , 50 , 54 ]. It is now well established that different types of carbohydrates can be oxidized at different rates in skeletal muscle due to the involvement of different transporter proteins that result in carbohydrate uptake [ 55 – 59 ]. Interestingly, combinations of glucose and sucrose or maltodextrin and fructose have been reported to promote greater exogenous rates of carbohydrate oxidation when compared to situations when single sources of carbohydrate are ingested [ 55 – 63 ]. These studies generally indicate a ratio of 1–1.2 for maltodextrin to 0.8–1.0 fructose seems to support the greatest rates of carbohydrate oxidation during exercise. Additional research on high molecular weight amylopectin indicates that there may be a benefit to the lower osmolality of the starch, allowing for greater consumption (100 g/hour) and possibly greater oxidation rates and performance improvement [ 64 – 67 ]. In addition to oxidation rates and carbohydrate types, the fasting status and duration of the exercise bout also function as key variables for athletes and coaches to consider. When considering duration, associated reviews have documented that bouts of moderate to intense exercise need to reach exercise durations that extend well into 90th minute of exercise before carbohydrate is shown to consistently yield an ergogenic outcome [ 41 , 68 , 69 ]. Of interest, however, not all studies indicate that shorter (60–75 min) bouts of higher intensity work may benefit from carbohydrate delivery. Currently the mechanisms surrounding these findings are, respectively, thought to be replacement of depleted carbohydrate stores during longer duration of moderate intensity while benefits seen during shorter, more intense exercise bouts are thought to operate in a central fashion. Moreover, these reviews have also pointed to the impact of fasting status on documentation of ergogenic outcomes [ 41 , 68 , 69 ]. In this respect, when studies require study participants to commence exercise in a fasted state, ergogenic outcomes are more consistently reported, yet other authors have questioned the ecological validity of this approach for competing athletes [ 43 ].

As it stands, the need for optimal carbohydrates in the diet for those athletes seeking maximal physical performance is unquestioned. Daily consumption of appropriate amounts of carbohydrate is the first and most important step for any competing athlete. As durations extend into 2 h, the need to deliver carbohydrate goes up, particularly when commencing exercise in a state of fasting or incomplete recovery. Once exercise ceases, several dietary strategies can be considered to maximally replace lost muscle and liver glycogen, particularly if a limited window of recovery exists. In these situations, the first priority should lie with achieving aggressive intakes of carbohydrate while strategies such as ingesting protein with lower carbohydrate amounts, carbohydrate and caffeine co-ingestion or certain forms of carbohydrate may also help to facilitate rapid assimilation of lost glycogen.

Considerable debate exists surrounding the amount of protein needed in an athlete’s diet [ 70 – 74 ]. Initially, it was recommended that athletes do not need to ingest more than the RDA for protein (i.e., 0.8 to 1.0 g/kg/d for children, adolescents and adults). However, research spanning the past 30 years has indicated that athletes engaged in intense training may benefit from ingesting about two times the RDA of protein in their diet (1.4–1.8 g/kg/d) to maintain protein balance [ 11 , 70 , 71 , 73 , 75 – 80 ]. If an insufficient amount of protein is consumed, an athlete will develop and maintain a negative nitrogen balance, indicating protein catabolism and slow recovery. Over time, this may lead to muscle wasting, injuries, illness, and training intolerance [ 76 , 77 , 81 ].

For people involved in a general fitness program or simply interested in optimizing their health, recent research suggests protein needs may also be above the RDA. Phillips and colleagues [ 76 ], Witard et al. [ 82 ], Jager et al. [ 11 ] and Tipton et al. [ 79 ] report that current evidence indicates optimal protein intakes in the range of 1.2–2.0 g/kg/day should be considered. In this respect, Morton and investigators [ 83 ] performed a meta-review and meta-regression involving 49 studies and 1863 participants and concluded that a daily protein intake of 1.62 g/kg/day may be an ideal place to start, with intakes beyond that providing no further contribution to increases in fat-free mass. In addition and in comparison to the RDA, non-exercising, older individuals (53–71 years) may also benefit from a higher daily protein intake (e.g., 1.0–1.2 g/kg/day of protein). Recent reports suggest that older muscle may be slower to respond and less sensitive to protein ingestion, typically requiring 40 g doses to robustly stimulate muscle protein synthesis [ 84 – 86 ]. Studies in younger individuals, however, have indicated that in the absence of exercise, a 20 g dose can maximize muscle protein synthesis [ 87 , 88 ] and if consumed after a multiple set workout consisting of several exercises that target large muscle groups a 40 g dose might be needed [ 89 ]. Consequently, it is recommended that athletes involved in moderate amounts of intense training consume 1.2–2.0 g/kg/day of protein (60–300 g/day for a 50–150 kg athlete) while athletes involved in high volume, intense training consume 1.7–2.2 g/kg/day of protein (85–330 g/day for a 50–150 kg athlete) [ 78 , 90 ]. This protein need would be equivalent to ingesting 3–15 three-ounce servings of chicken or fish per day for a 50–150 kg athlete [ 78 ]. Although smaller athletes typically can ingest this amount of protein, on a daily basis, in their normal diet, larger athletes often have difficulty consuming this much dietary protein. Additionally, a number of athletic populations are known to be susceptible to protein malnutrition (e.g., runners, cyclists, swimmers, triathletes, gymnasts, dancers, skaters, wrestlers, boxers, etc.) and consequently, additional counseling and education may be needed to help these athletes meet their daily protein needs. To this point, the periods of energy restriction to meet weight or aesthetic demands of their sports that are seemingly a part of the sport’s fabric creates an arguably greater need to understand that protein intake, quality and timing as well as combination with carbohydrate is particularly important to maintain lean body mass, training effects, and performance [ 25 ]. Overall, it goes without saying that care should be taken to ensure that athletes consume a sufficient amount of quality protein in their diet to maintain nitrogen balance.

Proteins differ based on their source, amino acid profile, and the methods of processing or isolating the protein undergoes [ 11 ]. These differences influence the availability of amino acids and peptides, which may possess biological activity (e.g., α-lactalbumin, ß-lactoglobulin, glycomacropeptides, immunoglobulins, lactoperoxidases, lactoferrin, etc.). Additionally, the rate of digestion and/or absorption and metabolic activity of the protein also are important considerations [ 91 ]. For example, different types of proteins (e.g., casein, whey, and soy) are digested at different rates, which may affect whole body catabolism and anabolism and acute stimulation of muscle protein synthesis (MPS) [ 91 – 96 ]. Therefore, care should be taken not only to make sure the athlete consumes enough protein in their diet but also that the protein is high quality. The best dietary sources of low fat, high quality protein are light skinless chicken, fish, egg whites, very lean cuts of beef and skim milk (casein and whey) while protein supplements routinely contain whey, casein, milk and egg protein. In what is still an emerging area of research, various plant sources of protein have been examined for their ability to stimulate increases in muscle protein synthesis [ 77 , 97 ] and promote exercise training adaptations [ 98 ]. While amino acid absorption from plant proteins is generally slower, leucine from rice protein has been found to be absorbed even faster than from whey [ 99 ], while digestive enzymes [ 100 ], probiotics [ 101 ] and HMB [ 102 ] can be used to overcome differences in protein quality. Preliminary findings suggest that rice [ 98 ] and pea protein [ 103 ] may be able to stimulate similar changes in fat-free mass and strength as whey protein, although the reader should understand that many other factors (dose provided, training status of participants, duration of training and supplementation, etc.) will ultimately impact these outcomes and consequently more research is needed.

While many reasons and scenarios exist for why an athlete may choose to supplement their diet with protein powders or other forms of protein supplements, this practice is not considered to be an absolute requirement for increased performance and adaptations. Due to nutritional, societal, emotional and psychological reasons, it is preferable for the majority of daily protein consumed by athletes to occur as part of a food or meal. However, we recognize and embrace the reality that situations commonly arise where efficiently delivering a high-quality source of protein takes precedence. Jager and colleagues [ 11 ] published an updated position statement of the International Society of Sports Nutrition that is summarized by the following points:

  • An acute exercise stimulus, particularly resistance exercise and protein ingestion both stimulate muscle protein synthesis (MPS) and are synergistic when protein consumption occurs before or after resistance exercise
  • For building and maintaining muscle mass, an overall daily protein intake of 1.4–2.0 g/kg/d is sufficient for most exercising individuals
  • Higher protein intakes (2.3–3.1 g/kg fat-free mass/d) may be needed to maximize the retention of lean body weight in resistance trained subjects during hypocaloric periods
  • Higher protein intakes (> 3.0 g protein/kg body weight/day) when combined with resistance exercise may have positive effects on body composition in resistance trained individuals (i.e., promote loss of fat mass)
  • Optimal doses for athletes to maximize MPS are mixed and are dependent upon age and recent resistance exercise stimuli. General recommendations are 0.25–0.55 g of a high-quality protein per kg of body weight, or an absolute dose of 20–40 g.
  • Acute protein doses should contain 700–3000 mg of leucine and/or a higher relative leucine content, in addition to a balanced array of the essential amino acids (EAAs)
  • Protein doses should ideally be evenly distributed, every 3–4 h, across the day
  • The optimal time period during which to ingest protein is likely a matter of individual tolerance; however, the anabolic effect of exercise is long-lasting (at least 24 h), but likely diminishes with increasing time post-exercise
  • Rapidly digested proteins that contain high proportions of EAAs and adequate leucine, are most effective in stimulating MPS
  • Different types and quality of protein can affect amino acid bioavailability following protein supplementation; complete protein sources deliver all required EAAs

The dietary recommendations of fat intake for athletes are similar to or slightly greater than dietary recommendations made to non-athletes to promote health. Maintenance of energy balance, replenishment of intramuscular triacylglycerol stores and adequate consumption of essential fatty acids are important for athletes, and all serve as reasons for an increased intake of dietary fat [ 104 ]. Depending upon the athlete’s training status or goals, the amount of dietary fat recommended for daily intake can change. For example, higher-fat diets appear to maintain circulating testosterone concentrations better than low-fat diets [ 105 – 107 ]. Additionally, higher fat intakes may provide valuable translational evidence to the documented testosterone suppression which can occur during volume-type overtraining [ 108 ]. Generally, it is recommended that athletes consume a moderate amount of fat (approximately 30% of their daily caloric intake), while proportions up to 50% of daily calories can be safely ingested by athletes during regular high-volume training [ 104 ]. In situations where an athlete may be interested in reducing their body fat, dietary fat intakes ranging from 0.5 to 1 g/kg/day have been recommended results in situations where daily fat intake might comprise as little as 20% of total calories in the diet [ 2 ]. This recommendation stems largely from available evidence in weight loss studies involving non-athletic individuals that people who are most successful in losing weight and maintaining the weight loss are those who ingest reduced amounts of fat in their diet [ 109 , 110 ] although this is not always the case [ 111 ]. Strategies to help athletes manage dietary fat intake include teaching them which foods contain various types of fat so that they can make better food choices and how to count fat grams [ 2 , 33 ].

For years, high-fat diets have been used by athletes with the majority of evidence showing no ergogenic benefit and consistent gastrointestinal challenges [ 112 ]. In recent years, significant debate has swirled regarding the impact of increasing dietary fat. One strategy, “train low, compete high”, refers to an acute pattern of dietary periodization whereby an athlete first follows a high-fat, low carbohydrate diet for one to 3 weeks while training before reintroducing carbohydrates back into the diet. While intramuscular adaptations result that may theoretically impact performance [ 113 , 114 ], no consistent, favorable impact on performance has been documented [ 112 , 115 ]. A variant of high-fat diets, ketogenic diets, have increased in popularity. While no exact prescription exists, nearly all ketogenic diet prescriptions derive at least 70–80% of their daily calories from dietary fat, prescribe a moderate amount of protein (20–25% total calories or 2.0–2.5 g/kg/day) and are largely devoid of carbohydrate (10–40 g per day). This diet prescription leads to a greater reliance on ketones as a fuel source. Currently, limited and mixed evidence remains regarding the overall efficacy of a ketogenic diet for athletes. In favor, Cox et al. [ 116 ] demonstrated that ketogenic dieting can improve exercise endurance by shifting fuel oxidation while Burke and colleagues [ 115 ] failed to show an increase in performance in a cohort of Olympic-caliber race walkers. Additionally, Jabekk and colleagues [ 117 ] reported decreases in body fat with no change in lean mass in overweight women who resistance trained for 10 weeks and followed a ketogenic diet. In light of the available evidence being limited and mixed, more human research needs to be completed before appropriate recommendations can be made towards the use of high fat diets for athletic performance.

Strategic eating and refueling

In addition to the general nutritional guidelines described above, research has also demonstrated that timing and composition of meals consumed may play a role in optimizing performance, training adaptations, and preventing overtraining [ 2 , 25 , 40 ]. In this regard, it takes about 4 h for carbohydrate to be digested and assimilated into muscle and liver tissues as glycogen. Consequently, pre-exercise meals should be consumed about four to 6 h before exercise [ 40 ]. This means that if an athlete trains in the afternoon, breakfast can be viewed to have great importance to top off muscle and liver glycogen levels. Research has also indicated that ingesting a light carbohydrate and protein snack 30 to 60 min prior to exercise (e.g., 50 g of carbohydrate and 5 to 10 g of protein) serves to increase carbohydrate availability toward the end of an intense exercise bout [ 118 , 119 ]. This also serves to increase availability of amino acids, decrease exercise-induced catabolism of protein, and minimize muscle damage [ 120 – 122 ]. Additionally, athletes who are going through periods of energy restriction to meet weight or aesthetic demands of sports should understand that protein intake, quality and timing as well as combination with carbohydrate is particularly important to maintain lean body mass, training effects, and performance [ 25 ]. When exercise lasts more than 1 h and especially as duration extends beyond 90 min, athletes should ingest glucose/electrolyte solutions (GES) to maintain blood glucose levels, prevent dehydration, and reduce the immunosuppressive effects of intense exercise [ 40 , 123 – 128 ]. Notably, this strategy becomes even more important if the athlete is under-fueled prior to the exercise task or is fasted vs. unfasted at the start of exercise [ 68 , 69 , 129 ]. Following intense exercise, athletes should consume carbohydrate and protein (e.g., 1 g/kg of carbohydrate and 0.5 g/kg of protein) within 30 min after exercise and consume a high carbohydrate meal within 2 h following exercise [ 2 , 74 ]. This nutritional strategy has been found to accelerate glycogen resynthesis as well as promote a more anabolic hormonal profile that may hasten recovery [ 120 , 130 , 131 ], but as mentioned above only when rapid glycogen restoration is needed or if the carbohydrate intake in the diet is adequate (< 6 g/kg/day) [ 53 , 132 ]. In other words, the total carbohydrate consumption and timing of carbohydrate consumption should be individualized to each athlete’s needs according to the goals of the training cycle and bout [ 112 ]. Finally, for two to 3 days prior to competition, athletes should taper training by 30 to 50% and consume an additional 200 to 300 g of carbohydrate each day in their diet. This eating strategy has been shown to supersaturate carbohydrate stores prior to competition and improve endurance exercise capacity [ 2 , 40 ]. Thus, the type of meal, amount of carbohydrate consumed, and timing of eating are important factors to maximize glycogen storage and in maintaining carbohydrate availability during training while also potentially decreasing the incidence of overtraining. The ISSN has adopted a position stand on nutrient timing in 2008 [ 133 ] that has been subsequently revised [ 13 ] and can be summarized with the following points:

  • Intramuscular and hepatic glycogen stores are best maximized by consumption of a high-carbohydrate diet (8–12 g/kg/day). Strategies such as aggressive carbohydrate feedings (~ 1.2 g/kg/hour) that favor high-glycemic (> 70) carbohydrates, addition of caffeine (3–8 mg/kg) and combining a moderate carbohydrate dose (0.8 g/kg/h) with protein (0.2–0.4 g/kg/h) have been shown to promote rapid restoration of glycogen stores.
  • High intensity (> 70% VO 2 Max) exercise bouts that extend beyond 90 min challenge fuel supply and fluid regulation. In these situations, it is advisable to consume carbohydrate at a rate of 30–60 g of carbohydrate/hour in a 6–8% carbohydrate-electrolyte solution (6–12 fluid ounces) every 10–15 min throughout the entire exercise bout. The importance of this strategy is increased when poor feeding or recovery strategies were employed prior to exercise commencement. Consequently, when carbohydrate delivery is inadequate, adding protein may help increase performance, mitigate muscle damage, promote euglycemia, and facilitate glycogen re-synthesis.
  • Consuming a diet that delivers adequate energy (minimum of 27–30 kcal/kg) and protein (1.6–1.8 g/kg/day), preferably with evenly spaced (every 3–4 h) protein feedings (0.25–0.40 g/kg/dose) during the day, should be considered for all exercising individuals.
  • Ingesting efficacious doses (10–12 g) of essential amino acids (EAAs) either in free form or as a protein bolus in 20–40 g doses (0.25–0.40 g/kg/dose) will maximally stimulate muscle protein synthesis (MPS).
  • Pre- and/or post-exercise nutritional interventions (carbohydrate + protein or protein alone) can be an effective strategy to support improvements in strength and body composition. However, the size (0.25–0.40 g/kg/dose) and timing (0–4 h) of a pre-exercise meal may impact the benefit derived from the post-exercise protein feeding.
  • Post-exercise ingestion (immediately-post to 2 h post) of high-quality protein sources stimulates robust increases in MPS. Similar increases in MPS have been found when high-quality proteins are ingested immediately before exercise.

Vitamins are essential organic compounds that serve to regulate metabolic and neurological processes, energy synthesis, and prevent destruction of cells. Fat-soluble vitamins include vitamins A, D, E, & K and the body stores fat-soluble vitamins in various tissues, which can result in toxicity if consumed in excessive amounts. Water-soluble vitamins consist of the entire complex of B-vitamins and vitamin C. Since these vitamins are water-soluble, excessive intake of these vitamins are eliminated in urine, with few exceptions (e.g. vitamin B6, which can cause peripheral nerve damage when consumed in excessive amounts). Table  1 describes the RDA, proposed ergogenic benefit, and summary of research findings for fat and water-soluble vitamins. Research has demonstrated that specific vitamins possess various health benefits (e.g., Vitamin E, niacin, folic acid, vitamin C, etc.), while few published studies have reported to find an ergogenic value of vitamins for athletes [ 134 – 138 ]. Alternatively, if an athlete is deficient in a vitamin, supplementation or diet modifications to improve vitamin status can consistently improve health and performance [ 139 ]. For example, Paschalis and colleagues [ 140 ] supplemented individuals who were low in vitamin C for 30 days and reported these individuals had significantly lower VO 2 Max levels than a group of males who were high in vitamin C. Further, after 30 days of supplementation, VO 2 Max significantly improved in the low vitamin C cohort as did baseline levels of oxidative stress of oxidative stress. Importantly, one must consider that some vitamins may help athletes tolerate training to a greater degree by reducing oxidative damage (Vitamin E, C) and/or help to maintain a healthy immune system during heavy training (Vitamin C). Alternatively, conflicting evidence has accumulated that ingesting high doses of Vitamins C and E may negatively impact intracellular adaptations seen in response to exercise training [ 141 – 144 ], which may consequently negatively impact an athlete’s performance. Furthermore, while optimal levels of vitamin D have been linked to improved muscle health [ 145 ] and strength [ 146 ] in general populations, research studies conducted in athletes generally fail to report on the ergogenic impact of vitamin D in athletes [ 147 , 148 ]. However, equivocal evidence from Wyon et al. [ 149 ] suggests vitamin D supplementation in elite ballet dancers improved strength and reduced risk for injuries. The remaining vitamins reviewed appear to have little ergogenic value for athletes who consume a normal, nutrient dense diet. Since dietary analyses of athletes commonly indicate that athletes fail to consume enough calories and subsequently may not be consuming adequate amounts of each vitamin, many sport dietitians and nutritionists recommend that athletes consume a low-dose daily multivitamin and/or a vitamin enriched post-workout carbohydrate/protein supplement during periods of heavy training [ 150 ]. Finally, athletes may desire to consume a vitamin or mineral for various health (non-performance) related reasons including niacin to elevate high density lipoprotein (HDL) cholesterol levels and decrease risk of heart disease (niacin), vitamin E for its antioxidant potential, vitamin D for its ability to preserve musculoskeletal function, or vitamin C to promote and maintain a healthy immune system.

Proposed nutritional ergogenic aids – vitamins

Recommended Dietary Allowances (RDA) based on the 1989 Food & Nutrition Board, National Academy of Sciences-National Research Council recommendations. Updated in 2001

Minerals are essential inorganic elements necessary for a host of metabolic processes. Minerals serve as structure for tissue, important components of enzymes and hormones, and regulators of metabolic and neural control. In athletic populations, some minerals have been found to be deficient while other minerals are reduced secondary to training and/or prolonged exercise. Notably, acute changes in sodium, potassium and magnesium throughout a continued bout of moderate to high intensity exercise are considerable. In these situations, athletes must work to ingest foods and fluids to replace these losses, while physiological adaptations to sweat composition and fluid retention will also occur to promote a necessary balance. Like vitamins, when mineral status is inadequate, exercise capacity may be reduced and when minerals are supplemented in deficient athletes, exercise capacity has been shown to improve [ 151 ]. However, scientific reports consistently fail to document a performance improvement due to mineral supplementation when vitamin and mineral status is adequate [ 134 , 152 , 153 ]. Table  2 describes minerals that have been purported to affect exercise capacity in athletes. Of the minerals reviewed, several appear to possess health and/or ergogenic value for athletes under certain conditions. For example, calcium supplementation in athletes susceptible to premature osteoporosis may help maintain bone mass [ 151 ]. For years, the importance of iron status in female athletes has been discussed [ 154 ] and more recent efforts have highlighted that iron supplementation in athletes prone to iron deficiencies and/or anaemia can improve exercise capacity [ 155 , 156 ]. Sodium phosphate loading can increase maximal oxygen uptake, anaerobic threshold, and improve endurance exercise capacity by 8 to 10% [ 157 ]. Increasing dietary availability of salt (sodium chloride) during the initial days of exercise training in the heat helps to maintain fluid balance and prevent dehydration. The American College of Sports Medicine (ACSM) recommendations for sodium levels (340 mg) represent the amount of sodium in less than 1/8 teaspoon of salt and recommended guidelines for sodium ingestion during exercise (300–600 mg per hour or 1.7–2.9 g of salt during a prolonged exercise bout) [ 158 – 161 ]. Finally, zinc supplementation during training can support changes in immune status in response to exercise training. Consequently, several minerals may enhance exercise capacity and/or training adaptations for athletes under certain conditions. However, there is little evidence that boron, chromium, magnesium, or vanadium affect exercise capacity or training adaptations in healthy individuals eating a normal diet. Sport nutritionists and dietitians should be aware of the specialized situations in which different types of minerals may provide support to bolster an athlete’s health or physical performance.

Proposed nutritional ergogenic aids – minerals

Recommended Dietary Allowances (RDA) based on the 2002 Food & Nutrition Board, National Academy of Sciences-National Research Council recommendations

a Estimated minimum requirement

The most important nutritional ergogenic aid for athletes is water and limiting dehydration during exercise is one of the most effective ways to maintain exercise capacity. Before starting exercise, it is highly recommended that individuals are adequately hydrated [ 162 ]. Exercise performance can be significantly impaired when 2% or more of body weight is lost through sweat (i.e., a 1.4 kg body weight loss from a 70-kg athlete). When one considers that average sweat rates are reported to be 0.5–2.0 L/hour during exercise and training [ 128 , 162 ], performance losses due to water loss can occur after just 60–90 min of exercise. Further, weight loss of more than 4% of body weight during exercise may lead to heat illness, heat exhaustion, heat stroke, and possibly death [ 128 ]. For this reason, it is critical that athletes adopt a mind set to prevent dehydration first by promoting optimal levels of pre-exercise hydration. Throughout the day and without any consideration of when exercise is occurring, a key goal is for an athlete to drink enough fluids to maintain their body weight. Next, athletes can promote optimal pre-exercise hydration by ingesting 500 mL of water or sports drinks the night before a competition, another 500 mL upon waking and then another 400–600 mL of cool water or sports drink 20–30 min before the onset of exercise. Once exercise commences, the athlete should strive to consume a sufficient amount of water and/or glucose-electrolyte solutions (i.e., sports drinks) during exercise to maintain hydration status. Consequently, to maintain fluid balance and prevent dehydration, athletes need to plan on ingesting 0.5 to 2 L/hour of fluid to offset weight loss. This requires frequent (every 5–15 min) ingestion of 12–16 fluid ounces of cold water or a sports drink during exercise [ 128 , 163 – 166 ]. Athletes should not depend on thirst to prompt them to drink because people do not typically get thirsty until they have lost a significant amount of fluid through sweat. Additionally, athletes should weigh themselves prior to and following exercise training to monitor changes in fluid balance and then can work to replace their lost fluid [ 128 , 163 – 166 ]. During and after exercise, athletes should consume three cups of water for every pound lost during exercise to promote adequate rehydration [ 128 ]. A primary goal soon after exercise should be to completely replace lost fluid and electrolytes during a training session or competition. Additionally, sodium intake in the form of glucose-electrolyte solutions (vs. only drinking water) and making food choices and modifications (added salt to foods) should be considered during the rehydration process to further promote euhydration [ 167 ]. Athletes should train themselves to tolerate drinking greater amounts of water during training and make sure that they consume more fluid in hotter/humid environments. Beyond nutrition, allowing one’s physiology the chance to acclimatize to the exercising environment for 10–14 days can help improve heat tolerance and promote thermoregulation. Finally, inappropriate and excessive weight loss techniques (e.g., cutting weight in saunas, wearing rubber suits, severe dieting, vomiting, using diuretics, etc.) are considered dangerous and should be prohibited. Sport nutritionists, dietitians, and athletic trainers can play an important role in educating athletes and coaches about proper hydration methods and supervising fluid intake during training and competition.

Dietary supplements and athletes

Educating athletes and coaches about nutrition and how to structure their diet to optimize performance and recovery are key areas of involvement for sport dietitians and nutritionists. Currently, use of dietary supplements by athletes and athletic populations is widespread while their overall need and efficacy of certain ingredients remain up for debate. Dietary supplements can play a meaningful role in helping athletes consume the proper amount of calories, macro- and micronutrients. Dietary supplements are not intended to replace a healthy diet. Numerous dietary ingredients have been investigated for potential benefits in an athletic population, to enhance training, recovery and/or performance. Supplementation with these nutrients in clinically validated amounts and at opportune times can help augment the normal diet to help optimize performance or support adaptations towards a training outcome. Sport dietitians and nutritionists must be aware of the current data regarding nutrition, exercise, and performance and be honest about educating their clients about results of various studies (whether pro or con). Currently, misleading information is available to the public and this position stand is intended to objectively rate many of the available ingredients. Additionally, athletes, coaches and trainers need to also heed the recommendations of scientists when recommendations are made according to the available literature and what will hopefully be free of bias. Throughout the next two sections of this paper, various nutritional supplements often taken by athletes will be categorized into three categories: Strong Evidence to Support Efficacy and Apparently Safe, Limited or Mixed Evidence to Support Efficacy, Little to No Evidence to Support Efficacy and/or Safety. Based on the available literature, the resulting classification and analysis focuses primarily on whether the proposed nutrient has been found to affect exercise and/or training adaptations through an increase in muscle hypertrophy and later for the supplement’s ergogenic potential. We recognize that some ingredients may exhibit little potential to stimulate training adaptations or operate in an ergogenic fashion, but may favorably impact muscle recovery or exhibit health benefits that may be helpful for some populations. These outcomes are not the primary focus of this review and consequently, will not be discussed with the same level of detail.

Convenience supplements

Convenience supplements are commonly found in the form of meal replacement powders (MRP’s), ready to drink supplements (RTD’s), energy bars, and energy gels. These products are typically fortified with vitamins and minerals and differ on the amount of carbohydrate, protein, and/or fat they contain. Uniqueness of these products come from the additional nutrients they contain that are purported to promote weight gain, alter body composition, enhance recovery, and/or improve performance. Most people view these supplements as a nutrient dense snack and/or use them to help control caloric intake when trying to gain and/or lose weight. MRP’s, RTD’s, and energy bars/gels can provide a convenient way for people to meet specific dietary needs and/or serve as good alternatives to fast food, foods of lower nutritional quality, and during times when travel or a busy schedule preclude the ability to consume fresh or other forms of whole food. Use of these types of products are particularly helpful in providing carbohydrate, protein, and other nutrients prior to and/or following exercise to optimize nutrient intake when an athlete doesn’t have time to sit down for a good meal or wants to minimize food volume. Consequently, meal replacements should be used in place of a meal during unique situations and are not intended to replace all meals. Care should also be taken to make sure they do not contain any banned or prohibited nutrients.

Muscle building supplements

The following section provides an analysis of the scientific literature regarding nutritional supplements purported to promote skeletal muscle accretion in conjunction with the completion of a well-designed exercise-training program. An overview of each supplement and a general interpretation of how they should be categorized is provided throughout the text. Table  3 summarizes how every supplement discussed in this article is categorized. However, within each category all supplements are ordered alphabetically. The reader is encouraged to consider that gains or losses in body masses may positively or negatively impact an individual’s athletic performance. For example, increases in body mass and lean mass are desired adaptations for many American football or rugby players and may improve performance in these activities. In contrast, decreases in body mass or fat mass may promote increases in performance such as cyclists and gymnasts whereby athletes such as wrestlers, weightlifters and boxers may need to rapidly reduce weight while maintaining muscle mass, strength and power.

Summary of categorization of dietary supplements based on available literature

Strong evidence to support efficacy and apparently safe

β-hydroxy β-methylbutyrate (hmb).

HMB is a metabolite of the amino acid leucine. It is well-documented that supplementing with 1.5 to 3 g/day of calcium HMB during resistance training can increase muscle mass (+ 0.5–1 kg greater than controls during 3–6 weeks of training) and strength particularly among untrained subjects initiating training [ 168 – 173 ] and the elderly [ 174 ]. The currently established minimal effective dose of HMB is 1.5 g per day, with 3 g per day offering additional benefits on lean body mass, while 6 g per day do not provide any additional gains in lean mass beyond what was reported with a 3 g dose [ 169 ]. To optimize HMB retention, its recommend to split the daily dose of 3 g into three equal doses of 1 g each (with breakfast, lunch or pre-exercise, bedtime) [ 174 ]. From a safety perspective, dosages of 1.5–6 g per day have been well tolerated [ 15 , 169 , 170 ]. The effects of HMB supplementation in trained athletes are less clear with selected studies reporting non-significant gains in muscle mass [ 175 – 177 ]. In this respect, it has been suggested by Wilson and colleagues [ 15 ] that program design (periodized resistance training models) and duration of supplementation (minimum of 6 weeks) likely operate as key factors. In 2015, Durkalec-Michalski and investigators [ 178 ] supplemented highly trained rowers ( n  = 16) in a randomized, double-blind, crossover fashion with either 3 g per day of calcium-HMB or a placebo. Before and after each supplementation period, body composition and performance parameters were assessed. When HMB was provided, fat mass was significantly reduced while changes in lean mass were not significant between groups. The same research group published data of 58 highly trained males athletes who supplemented with either 3 g of calcium-HMB or placebo for 12 weeks in a randomized, double-blind, crossover fashion [ 179 ]. In this report, fat mass was found to be significantly reduced while fat-free mass was significantly increased. Finally, Durkalec-Michalski and investigators [ 180 ] supplemented 42 highly-trained combat sport athletes for 12 weeks with either a placebo or 3 g of calcium-HMB in a randomized, double-blind, crossover fashion. When HMB was provided, fat-free mass was shown to increase ( p  = 0.049) while fat mass was significantly reduced in comparison to the changes seen when placebo was provided. In conclusion, a growing body of literature continues to offer support that HMB supplementation at dosages of 1.5–3 g for durations as short as three to 4 weeks in untrained populations and longer durations (12 weeks) in trained populations can lead to improvements in fat mass and fat-free mass while participating in various forms of exercise training.

Creatine monohydrate

In our view, the most effective nutritional supplement available to athletes to increase high intensity exercise capacity and muscle mass during training is creatine monohydrate. Numerous studies have indicated that creatine supplementation increases body mass and/or muscle mass during training [ 181 , 182 ]. Body mass increases are typically one to two kilograms greater than controls during 4–12 weeks of training [ 182 ]. The gains in muscle mass appear to be a result of an improved ability to perform high intensity exercise enabling an athlete to train harder and thereby promote greater training adaptations and muscle hypertrophy [ 183 – 186 ]. The only clinically significant side effect occasionally reported from creatine monohydrate supplementation has been the potential for weight gain [ 181 , 182 , 187 , 188 ]. Although concerns have been raised about the safety and possible side effects of creatine supplementation [ 189 , 190 ], multiple shorter [ 191 – 193 ] and long-term safety studies have reported no apparent side effects [ 188 , 194 , 195 ] and/or that creatine monohydrate may lessen the incidence of injury during training [ 196 – 199 ]. Consequently, supplementing the diet with creatine monohydrate and/or creatine containing formulations seems to be a safe and effective method to increase muscle mass. The ISSN position stand on creatine monohydrate [ 10 ] summarizes their findings as this:

  • Creatine monohydrate is the most effective ergogenic nutritional supplement currently available to athletes in terms of increasing high-intensity exercise capacity and lean body mass during training.
  • Creatine monohydrate supplementation is not only safe, but has been reported to have a number of therapeutic benefits in healthy and diseased populations ranging from infants to the elderly. There is no compelling scientific evidence that the short- or long-term use of creatine monohydrate (up to 30 g/day for 5 years) has any detrimental effects on otherwise healthy individuals or among clinical populations who may benefit from creatine supplementation.
  • If proper precautions and supervision are provided, creatine monohydrate supplementation in children and adolescent athletes is acceptable and may provide a nutritional alternative with a favorable safety profile to potentially dangerous anabolic androgenic drugs. However, it is recommended that creatine supplementation only be considered for use by younger athletes who: a) are involved in serious/competitive supervised training; b) are consuming a well-balanced and performance enhancing diet; c) are knowledgeable about the appropriate use of creatine; and d) do not exceed recommended dosages.
  • Label advisories on creatine products that caution against usage by those under 18 years old, while perhaps intended to insulate their manufacturers from legal liability, are likely unnecessary given the science supporting creatine’s safety, including in children and adolescents.
  • At present, creatine monohydrate is the most extensively studied and clinically effective form of creatine for use in nutritional supplements in terms of muscle uptake and ability to increase high-intensity exercise capacity.
  • The addition of carbohydrate or carbohydrate and protein to a creatine supplement appears to increase muscular uptake of creatine, although the effect on performance measures may not be greater than using creatine monohydrate alone.
  • The quickest method of increasing muscle creatine stores appears to be to consume ~ 0.3 g/kg/day of creatine monohydrate for 5–7 days followed by 3–5 g/day thereafter to maintain elevated stores. Initially, ingesting smaller amounts of creatine monohydrate (e.g., 3–5 g/day) will increase muscle creatine stores over a three to 4 week period, however, the initial performance effects of this method of supplementation are less supported.
  • Clinical populations have been supplemented with high levels of creatine monohydrate (0.3–0.8 g/kg/day equivalent to 21–56 g/day for a 70-kg individual) for years with no clinically significant or serious adverse events.
  • Further research is warranted to examine the potential medical benefits of creatine monohydrate and precursors like guanidinoacetic acid on sport, health and medicine.

Essential amino acids (EAA)

Research examining the impact of the essential amino acids on stimulating muscle protein synthesis is an extremely popular area. Collectively, this data indicates that ingesting 6–12 g of the essential amino acids (EAA) in the absence of feeding [ 200 ] and prior to [ 201 , 202 ] and/or following resistance exercise stimulates protein synthesis [ 202 – 208 ], with this response being largely independent of the protein source or food type [ 209 ]. Theoretically, this may enhance increases in fat-free mass, but to date limited evidence exists to demonstrate that supplementation with non-intact sources of EAAs (e.g., free form amino acids) while resistance training positively impacts fat-free mass accretion. Moreover, other research has indicated that changes in muscle protein synthesis may not correlate with phenotypic adaptations to exercise training [ 210 ]. An abundance of evidence is available, however, to indicate that ingestion of high-quality protein sources can heighten adaptations to resistance training [ 211 ]. While various methods of protein quality assessment exist, most of these approaches center upon the amount of EAAs that are found within the protein source, and in nearly all situations, the highest quality protein sources are those containing the highest amounts of EAAs. To this point, a number of published studies are available that state the EAAs operate as a prerequisite to stimulate peak rates of muscle protein synthesis [ 212 – 215 ]. To better understand the impact of ingesting free-form amino acids versus an intact protein source, Katsanos et al. [ 216 ] administered similar doses of the essential amino acids (6.72 g) as part of an intact protein (15 g of whey protein isolate) source or as free amino acids while completing a resistance training program in elderly adults. Protein accrual was greater when the amino acid dose was provided in an intact source. While the age of the participants in this study may have impacted outcomes [ 217 ], this study’s results do highlight the need for more research to better understand to what extent training adaptations are due to the EAA content or if additional benefits are present from ingesting an intact protein source.

While the EAAs are comprised of nine separate amino acids, some individual EAAs have received considerable attention for their potential role in impacting protein translation and muscle protein synthesis. In this respect, the branched-chain amino acids have been highlighted for their predominant role in stimulating muscle protein synthesis [ 218 , 219 ]. To this point, Karlsson and colleagues [ 220 ] demonstrated significantly higher increases in p70s6k expression in recovery from a single bout of lower-body resistance exercise in seven male participants after ingesting a BCAA solution containing 100 mg/kg BCAA when compared to ingesting a placebo. Interestingly, Moberg and investigators [ 221 ] had trained volunteers complete a standardized bout of resistance training in conjunction with ingestion of placebo, leucine, BCAA or EAA while measuring changes in post-exercise activation of p70s6k. They concluded that EAA ingestion led to a nine-fold greater increase in p70s6k activation and that these results were primarily attributable to the BCAAs. Finally, a 2017 study by Jackman et al. [ 222 ] compared the ability of a 5.6 g dose of BCAAs (versus a placebo) to stimulate increases in muscle protein synthesis. Myofibrillar muscle protein synthesis rates were increased significantly (~ 20%, p  < 0.05) in comparison to a placebo. While significant, this magnitude of change was notably less than the post-exercise MPS responses seen when doses of whey protein that delivered similar amounts of the BCAAs were consumed [ 88 , 223 ]. These outcomes led the authors to conclude that the full complement of EAAs was advised to maximally stimulate increases in MPS.

Of all the interest captured by the BCAAs, leucine is accepted to be the primary driver of acute changes in protein translation. In this respect, Dreyer et al. [ 224 ] and others [ 225 ] have reported that providing leucine after completion of resistance exercise can further potentiate increases in mTOR signalling and protein translation. In this respect, Jager et al. [ 11 ] have highlighted that an ideal dose of leucine to stimulate increases in protein translation is likely somewhere between 1.7–3.5 g.

A growing body of literature is available that suggests higher amounts of protein are needed by exercising individuals to optimize exercise training adaptations [ 11 , 83 , 211 , 226 ]. Collectively, these sources indicate that people undergoing intense training with the primary intention to promote accretion of fat-free mass should consume between 1.4–2.0 g of protein per kilogram of body weight per day [ 83 , 226 ]. Tang and colleagues [ 95 ] conducted a classic study that examined the ability of three different sources of protein (hydrolyzed whey isolate, micellar casein and soy isolate) to stimulate acute changes in muscle protein synthesis both at rest and after a single bout of resistance exercise. These authors concluded that all three protein sources significantly increased muscle protein synthesis rates both at rest and in response to resistance exercise. When this response is extrapolated over the course of several weeks, multiple studies have reported on the ability of different forms of protein to significantly increase fat-free mass while resistance training [ 70 , 227 – 232 ]. Cermak et al. [ 211 ] performed a meta-analysis that examined the impact of protein supplementation on changes in strength and fat-free mass. Data from 22 separate published studies that included 680 research participants were included in the analysis. These authors concluded that protein supplementation demonstrated a positive effect of fat-free mass and lower-body strength in both younger and older participants. Similarly, Morton and investigators [ 83 ] published results from a meta-analysis that also included a meta-regression approach involving data from 49 studies and 1863 participants. They concluded that the ability of protein to positively impact fat-free mass accretion increases up to approximately 1.62 g of protein per kilogram of body weight per day whereby higher amounts beyond that do not appear to promote greater gains in fat-free mass. Although more research is necessary in this area, evidence clearly indicates that protein needs of individuals engaged in intense training are elevated and consequently those athletes who achieve higher intakes of protein while training promote greater changes in fat-free mass. Beyond the impact of protein to foster greater training-induced adaptations such as increases in strength and muscle mass, several studies have examined the ability of different types of protein to stimulate changes in fat-free mass [ 229 , 231 , 233 – 235 ] while several studies and reviews have critically explored the role protein may play in achieving weight loss in athletes [ 236 , 237 ] as well as during periods of caloric restriction [ 238 , 239 ]. Therefore, it is simplistic and misleading to suggest that there is no data supporting contentions that athletes need more protein in their diet and/or there is no potential ergogenic value of incorporating different types of protein into the diet. It is the position stand of ISSN that exercising individuals need approximately 1.4 to 2.0 g of protein per kilogram of bodyweight per day [ 11 ].

Limited or mixed evidence to support efficacy

Adenosine − 5′-triphosphate (atp).

ATP is the primary intracellular energy source and in addition, has extensive extracellular functions including the increase in skeletal muscle calcium permeability and vasodilation. While intravenous administration of ATP is bioavailable [ 240 ], several studies have shown that oral ATP is not systematically bioavailable [ 241 ]. However, chronic supplementation with ATP increases the capacity to synthesize ATP within the erythrocytes without increasing resting concentrations in the plasma, thereby minimizing exercise-induced drops in ATP levels [ 242 ]. Oral ATP supplementation has demonstrated initial ergogenic properties, after a single dose, improving total weight lifted and total number of repetitions [ 243 ]. ATP may increase blood flow to the exercising muscle [ 244 ] and may reduce fatigue and increase peak power output during later bouts of repeated bouts exercise [ 242 ]. ATP may also support greater recovery and lean mass maintenance under high volume training [ 245 ], however, this has only been reported in one previous study. In addition, ATP supplementation in clinical populations has been shown to improve strength, reduce pain after knee surgery, and reduce the length of the hospital stay [ 246 ]. However, given the limited number of human studies of ATP on increasing exercise-induced gains in muscle mass, more chronic human training studies are warranted.

Branched chain amino acids (BCAA)

BCAA supplementation has been reported to decrease exercise-induced protein degradation and/or muscle enzyme release (an indicator of muscle damage) possibly by promoting an anti-catabolic hormonal profile [ 118 , 247 , 248 ] and more recent studies support their ability to favorably promote responses to damaging eccentric muscle contractions [ 249 , 250 ]. Leucine, in particular, is recognized as a keystone of sorts that when provided in the correct amounts (3–6 g) activates the mTORC1 complex resulting in favorable initiation of translation [ 251 ]. To highlight this impact for leucine, varying doses of whey protein and leucine levels were provided to exercising men at rest and in response to an acute bout of lower-body resistance exercise to examine the muscle protein synthetic response. Interestingly, when a low dose of whey protein (6.25 g) was enriched with leucine to equal the leucine content found in a 25-g dose of whey protein, the ability to stimulate muscle protein synthesis was retained. While the 25-g dose of whey protein did favorably sustain the increases in muscle protein synthesis, the added leucine highlights an important role for leucine in stimulating muscle protein synthesis in response to resistance exercise [ 223 ]. For these reasons, it has been speculated that the leucine content of whey protein and other high-quality protein sources have been suggested to be primary reasons for their ability to stimulate favorable adaptations to resistance training [ 252 , 253 ]. Theoretically, BCAA supplementation during intense training may help minimize protein degradation and thereby lead to greater gains in (or limit losses of) fat-free mass, but only limited evidence exists to support this hypothesis. For example, Schena and colleagues [ 254 ] reported that BCAA supplementation (~ 10 g/d) during 21-days of trekking at altitude increased fat free mass (1.5%) while subjects ingesting a placebo had no change in muscle mass. Bigard and associates [ 255 ] reported that BCAA supplementation appeared to minimize loss of muscle mass in subjects training at altitude for 6 weeks. Finally, Candeloro and coworkers [ 256 ] reported that 30 days of BCAA supplementation (14 g/day) promoted a significant increase in muscle mass (1.3%) and grip strength (+ 8.1%) in untrained subjects. Alternatively, Spillane and colleagues [ 257 ] reported that 8 weeks of resistance training while supplementing with either 9 g of BCAAs or placebo did not impact body composition or muscle performance. Most recently, Jackman et al. [ 222 ] examined the ability of an acute dose of branched-chain amino acids to stimulate increases in muscle protein synthesis. While acute ingestion of BCAAs did promote a 22% greater increase in muscle protein synthesis when compared to a placebo, the determined rates were 50% lower than what is commonly seen when a dose of whey protein containing similar amounts of BCAAs is ingested. As mixed outcomes cloud the ability to make clear determinations, studies strongly suggest a mechanistic role for BCAAs and in particular leucine, yet translational data fails to consistently support the need for BCAA supplementation. Alternatively, multiple studies do support BCAAs ability to mitigate recovery from damaging exercise while their ability to favorably impact resistance training adaptations needs further research. This will be discussed in a later section.

Phosphatidic acid

Phosphatidic acid (PA) is a diacyl-glycerophospholipid that is enriched in eukaryotic cell membranes and it can act as a signalling lipid [ 258 ]. Interestingly, PA has been repeatedly shown to activate the mammalian target of rapamycin (mTOR) signalling in muscle; an effect which ultimately leads to increases in muscle protein synthesis. For instance, Fang et al. [ 259 ] demonstrated that PA activates mTOR in vitro. Hornberger et al. [ 260 ] also reported that mechanical stretching of skeletal muscle in situ promotes an increase in intramuscular PA levels and this effect was associated with the activation of mTOR signalling. To date, two chronic human supplementation studies have been performed whereby PA supplementation (750 mg/day) occurred in subjects engaged in resistance training. Hoffman et al. [ 261 ] reported that PA supplementation increased whole-body lean body mass (LBM) by 1.7 kg, whereas the placebo group demonstrated no relative change in LBM (0.1 kg; p  = 0.065 between groups). Joy et al. [ 262 ] performed a similar eight-week study with more participants and supervised training sessions, and reported that PA supplementation significantly increased LBM by 2.4 kg, whereas the placebo group demonstrated marginal increases in LBM (1.2 kg; p  < 0.05 between groups). A third study confirmed the beneficial effects of PA on exercise-induced gains in lean body mass [ 263 ]. The currently established dose of PA is 750 mg per day and another study investigating lower doses, 375 and 250 mg per day, failed to show significant benefits on lean body mass [ 264 ]. Hence, preliminary human research suggests that PA supplementation can increase anabolic signalling in skeletal muscle and enhance gains in muscle mass with resistance training. Given that PA supplementation studies are in their infancy relative to other muscle-building supplements (e.g., whey protein, creatine, HMB, etc.), future studies are needed in order to determine the optimal dosage, timing, and duration of supplementation needed for optimal muscle mass gains.

Little to no evidence to support efficacy and/or safety

Agmatine sulfate.

Agmatine, the decarboxylation product of the amino acid L-arginine, has shown different biological effects in different in vitro and animal models [ 265 ] indicating potential benefits in an athletic population. Agmatine is thought to improve insulin release and glucose uptake, assist in the secretion of luteinizing hormone, influence the nitric oxide signalling pathway, offer protection from oxidative stress, and is potentially involved in neurotransmission [ 266 ]. It is mostly found in fermented foods [ 267 ], with higher levels found in alcoholic beverages. Currently, nearly all research involving agmatine is commonly from animal research models and no human studies have been conducted to examine its impact on blood flow or impacting resistance training adaptations such as strength and body composition. There does not appear to be any scientific evidence that Agmatine supports increases in lean body mass or muscular performance.

α-ketoglutarate (α-KG)

α-ketoglutarate (α-KG) is an intermediate in the Krebs cycle that is involved in aerobic energy metabolism and may function to stimulate nitric oxide production. There is some clinical evidence that α-KG may serve as an anticatabolic nutrient after surgery [ 268 , 269 ]. However, it is unclear whether α-KG supplementation during training may affect training adaptations. Very little research has been conducted on just alpha-ketoglutarate in humans to examine exercise outcomes. For example, Little and colleagues [ 270 ] supplemented with creatine, a combination of creatine, α-KG, taurine, BCAA and medium-chain triglycerides, or a placebo. The combination of nutrients increased the maximal number of bench press repetitions completed and Wingate peak power while no changes were reported in the placebo group. Campbell and investigators [ 271 ] supplemented 35 healthy trained men with 2 g of arginine and 2 g of α-KG or placebo in a double-blind manner while resistance training for 8 weeks. Supplementation with arginine + α-KG increased bench press strength and Wingate peak power, but did not impact body composition. Finally, Willoughby and colleagues [ 272 ] examined the results of arginine α-KG supplementation in relation to increasing nitric oxide production (vasodilation during resistance exercise), hemodynamics, brachial artery flow, circulating levels of l-arginine, and asymmetric dimethyl arginine in active males. This study found that although plasma L-arginine increased, there was no significant impact of supplementation on nitric oxide production after a bout of resistance exercise. Due to the lack of research on α-KG examining its impact on exercise training adaptations, its use cannot be recommended at this time.

Arginine is commonly classified as a conditionally essential amino acid and has been linked to nitric oxide production and increases in blood flow that are purported to then stimulate enhanced nutrient and hormone delivery and favorably impact resistance training adaptations [ 273 ]. To date, few studies have examined the independent impact of arginine on the ability to enhance fat-free mass increases while resistance training. Tang and colleagues [ 274 ] used an acute model to examine the ability of an oral 10-g dose of arginine to stimulate changes in muscle protein synthesis. These authors reported that arginine administration failed to impact muscle protein synthesis or femoral artery blood flow. Growth hormone levels did rise in response to arginine ingestion, which contrasts with the findings of Forbes et al., [ 275 ] who reported a blunting of growth hormone production after acute ingestion of arginine in strength trained males. Regardless, the Tang study [ 274 ] and others [ 276 , 277 ] failed to link the increase in growth hormone to changes in rates of muscle protein synthesis. Notably, other studies have also failed to show a change in blood flow after arginine ingestion, one of its key purported benefits [ 272 , 278 ]. Campbell and colleagues published outcomes from an 8 week resistance training study that supplemented healthy men in a double-blind fashion with either a placebo or 2 g of arginine and 2 g of α-ketoglutarate. No changes in fat mass or fat-free mass were reported in this study. Therefore, due to the limited data of arginine supplementation on stimulating further increases of exercise in muscle mass, its use for is not recommended at this time.

Boron is a trace mineral whose physiological role is not clearly understood. A number of proposed functions have been touted for boron: vitamin D metabolism, macromineral metabolism, immune support, increase testosterone levels and promote anabolism [ 279 ]. Due to a lack of scientific evidence surrounding boron, no official Daily Reference Intake (DRI) is established. Several studies have evaluated the effects of boron supplementation during training on strength and body composition alterations. However, these studies (conducted on male bodybuilders) indicate that boron supplementation (2.5 mg/d) had no significant impact on muscle mass or strength [ 280 , 281 ]. Further, two investigations [ 282 , 283 ] examined the impact of boron supplementation on bone mineral density in athletic and sedentary populations. In both investigations, boron supplementation did not significantly influence bone mineral density. Therefore, due to the limited findings on boron supplementation, its use is not recommended, and more research is warranted to determine its physiological impact.

Chromium is a trace mineral that is actively involved in macronutrient metabolism. Clinical studies have suggested that chromium potentiates the effects of insulin, particularly in diabetic populations. Due to its close interaction with insulin, chromium supplementation has been theorized to impact anabolism and exercise training adaptations. Initial research was promising with chromium supplementation being associated with increases in muscle and strength, particularly in women [ 284 – 286 ]. Subsequent well-controlled research studies [ 287 ] since that time have consistently failed to report a benefit for chromium supplementation (200–800 μg/d for 4–16 weeks) [ 288 – 294 ]. Most recently, chromium supplementation was investigated for its ability to impact glycogen synthesis after high-intensity exercise and was found to exert no impact over recovery of glycogen [ 295 ]. In summary, chromium supplementation appears to exert very little potential for its ability to stimulate or support improvements in fat-free mass. Therefore, due to the limited data of arginine supplementation on stimulating further increases of exercise in muscle mass, its use for is not recommended at this time.

Conjugated linoleic acids (CLA)

Animal studies indicate that adding CLA to dietary feed decreases body fat, increases muscle and bone mass, has anti-cancer properties, enhances immunity, and inhibits progression of heart disease [ 296 – 298 ]. Although animal studies are impressive [ 299 – 301 ], human studies, at best, suggest a modest ability, independent of exercise or diet changes, of CLA to stimulate fat loss [ 302 – 305 ]. Moreover, very little research has been conducted on CLA to better understand if any scenario exists where its use may be justified. Initial work by Pinkoski et al. [ 306 ] suggested that CLA supplementation may help minimize catabolism while resistance training, but overall improvements in body composition from this study failed to yield positive outcomes. Two studies are available that supplemented exercising younger [ 307 ] and older individuals [ 308 ] with a combination of CLA and creatine and reported significant improvements in strength and body composition, but these results are thought to be the result of creatine. Currently, it seems there is little evidence that CLA supplementation during training can affect lean tissue accretion and has limited efficacy [ 309 ].

D-aspartic acid

Also known as aspartate, aspartic acid is a non-essential amino acid. Two isomers exist within aspartic acid: L-Aspartic acid and D-Aspartic acid. D-Aspartic acid is thought to help boost athletic performance and function as a testosterone booster. It is also used to conserve muscle mass. While limited research is available in humans examining D-aspartic, Willoughby and Leutholtz [ 310 ] published a study to determine the impact of D-aspartic acid in relation to testosterone levels and performance in resistance-trained males. The results showed D-aspartic acid did not impact testosterone levels nor did it improve any aspect of performance. In agreement, Melville and colleagues [ 311 ] had participants supplement with either three or 6 g of D-aspartic acid and concluded that neither dose of D-aspartic acid stimulated any changes in testosterone and other anabolic hormones. Later, Melville et al. [ 312 ] supplemented 22 men in a randomized, double-blind fashion with either a placebo or 6 g of D-aspartic acid and concluded that 12 weeks of supplementation exerted no impact on resting levels of free or total testosterone and all changes observed in strength or hypertrophy were similar to what was experienced in the placebo group. Based on the currently available literature, D-aspartic acid is not recommended to improve muscle health.

Ecdysterones

Ecdysterones (also known as ectysterone, 20 β-Hydroxyecdysterone, turkesterone, ponasterone, ecdysone, or ecdystene) are naturally derived phytoecdysteroids (i.e., insect hormones). They are typically extracted from the herbs Leuza rhaptonticum sp., Rhaponticum carthamoides , or Cyanotis vaga . They can also be found in high concentrations in the herb Suma (also known as Brazilian Ginseng or Pfaffia). Initial interest was generated for ecdysterones due to reports of research from Russia and Czechoslovakia that indicated a potential physiological benefit in insects and animals [ 313 – 316 ]. A review by Bucci on various herbals and exercise performance also mentioned suma (ecdysterone) [ 317 ]. Unfortunately, the initial work was available in obscure journals with sub-standard study designs and presentation of results. In 2006, Wilborn and coworkers [ 318 ] completed what remains as the only study in humans to examine the impact of ecdysterones while resistance training. Herein, a 200 mg daily dose of 20-hydroxyecdysone over 8 weeks yielded no impact on changes in fat free mass or anabolic/catabolic hormone status. Ecdysterones are not recommended for supplementation to increase training adaptations or performance.

Fenugreek extract

Fenugreek ( trigonella foenum-graecum ) is an Ayurvedic herb historically used to enhance masculinity and libido. Fenugreek extract has been shown to increase testosterone levels by decreasing the activity of the aromatase enzyme metabolizing testosterone into estradiol [ 319 , 320 ]. Initial research by Poole et al. [ 321 ] supplemented resistance trained men in a randomized, double-blind fashion with a placebo or 500 mg of Fenugreek extract. After 8 weeks of supplementing and resistance training, significantly greater improvements in body fat, lower body strength, and upper body strength were observed. Wankhede and colleagues [ 320 ] reported a significant increase in repetitions performed to failure using the bench press and a reduction in body fat when 600 mg Fenugreek extract was consumed while following a resistance training program. Initial research using Fenugreek extract suggests it may help improve resistance-training adaptations, but more research in different populations is needed before any further recommendations can be made.

Gamma oryzanol (ferulic acid)

Gamma oryzanol is a mixture of a plant sterol and ferulic acid theorized to increase anabolic hormonal responses, strength and muscle mass during training [ 322 , 323 ]. Although data are limited, one study reported no effect of 0.5 g/d of gamma oryzanol supplementation on strength, muscle mass, or anabolic hormonal profiles during 9 weeks of training [ 324 ]. Most recently, Eslami and colleagues [ 325 ] supplemented healthy male volunteers with either gamma oryzanol or placebo for 9 weeks while resistance training. In this study, changes in body composition were not realized, but a significant increase in strength was found in the bench press and leg curl exercise. With limited research of mixed outcomes at this point, no conclusive recommendation can be made at this time as more research is needed to fully determine what impact, if any, gamma oryzanol supplementation may have in exercising individuals.

Glutamine is the most plentiful non-essential amino acid in the body and plays several important physiological roles [ 74 , 326 , 327 ]. Glutamine has been reported to increase cell volume and stimulate protein [ 328 – 330 ] and glycogen synthesis [ 331 ]. Despite its important role in physiological processes, there is no compelling evidence to support the use of glutamine supplementation in terms of increasing lean body mass and a 2008 review by Gleeson concluded that minimal evidence is available to support glutamine’s purported role in exercise and sport training [ 332 ]. Initial research by Colker and associates [ 333 ] reported that subjects who supplemented their diet with glutamine (5 g) and BCAA (3 g) enriched whey protein (40 g) during resistance training promoted about a two pound greater gain in muscle mass and greater gains in strength than ingesting whey protein alone. In contrast, Kerksick and colleagues [ 232 ] reported no additional impact on strength, endurance, body composition and anaerobic power of combining 5 g of glutamine and 3 g of BCAAs to 40 g of whey protein in healthy men and women who resistance trained for 10 weeks. In addition, Antonio et al. [ 334 ] reported that high-dose glutamine ingestion (0.3 g/kg) offered no impact of the number of repetitions completed using the leg press or bench press exercises. In a well-designed investigation, Candow and co-workers [ 335 ] studied the effects of oral glutamine supplementation combined with resistance training in young adults. Thirty-one participants were randomly allocated to receive either glutamine (0.9 g/kg of lean tissue mass) or a maltodextrin placebo (0.9 g/kg of lean tissue mass) during 6 weeks of total body resistance training. The authors concluded glutamine supplementation during resistance training had no significant effect on muscle performance, body composition or muscle protein degradation in young healthy adults. While there may be other beneficial uses for glutamine supplementation (i.e. gastrointestinal health and peptide uptake in stressed populations [ 336 ] and, as mentioned previously, mitigation of soreness and recovery of lost force production [ 337 ]), there does not appear to be any scientific evidence that it supports increases in lean body mass or muscular performance.

Growth hormone releasing peptides (GHRP) and secretagogues

Growth hormone releasing peptides (GHRP) and other non-peptide compounds (secretagogues) facilitate growth hormone (GH) release [ 338 , 339 ], and can impact sleep patterns, food intake and cardiovascular functioning [ 340 ] along with improvements in lean mass in clinical wasting states [ 341 ]. These observations have served as the basis for development of nutritionally-based GH stimulators (e.g., amino acids, pituitary peptides, “pituitary substances”, Macuna pruriens , broad bean, alpha-GPC, etc.) and continue to capture interest by sporting populations for their potential to impact growth hormone secretion, recovery and robustness of training [ 342 ]. Although there is clinical evidence that pharmaceutical grade GHRP’s and some non-peptide secretagogues can increase GH and IGF-1 levels at rest and in response to exercise, it has not been demonstrated that such increases lead to an increase in skeletal muscle mass [ 343 ]. Finally, Chromiak and Antonio [ 344 ] reported that oral ingestion of many secretagogues fail to consistently stimulate hormone increases in growth hormone and fail to stimulate greater changes in muscle mass or strength. Currently, there is no convincing scientific evidence that secretagogues support increases in lean body mass or muscular performance.

Isoflavones

Isoflavones are naturally occurring non-steroidal phytoestrogens that have a similar chemical structure as ipriflavone (a synthetic flavonoid drug used in the treatment of osteoporosis) [ 345 – 347 ]. For this reason, soy protein (which is an excellent source of isoflavones) and isoflavone extracts have been investigated in the possible treatment of osteoporosis as well as their role in body composition changes and changes in cardiovascular health markers. In this respect, multiple studies have supported the ability of isoflavone supplementation in older women alone [ 348 ] and in combination with exercise over the course of 6–12 months to improve various body composition parameters [ 349 – 351 ]. Findings from these studies have some applications to sedentary, postmenopausal women. However, there are currently no peer-reviewed data indicating that isoflavone supplementation affects exercise, body composition, or training adaptations in physically active individuals. For example, Wilborn and colleagues [ 318 ] reported that 8 weeks of supplementing with isoflavones with resistance training did not significantly impact strength or body composition.

Ornithine-α-ketoglutarate (OKG)

OKG (via enteral feeding) has been shown to significantly shorten wound healing time and improve nitrogen balance in severe burn patients [ 352 , 353 ]. A 2004 review by Cynober postulated that OKG may operate as a precursor to arginine and nitric oxide, but the overall lack of efficacy for arginine and other precursors limits the potential of OKG. Because of its ability to improve nitrogen balance, OKG may provide some value for athletes engaged in intense training. A study by Chetlin and colleagues [ 354 ] reported that OKG supplementation (10 g/day) during 6 weeks of resistance training significantly increased upper body strength. However, no significant differences were observed in lower body strength, training volume, gains in muscle mass, or fasting insulin and growth hormone. Since the previously published version of this review, no additional human research has been published and consequently, no further recommendations can be made regarding OKG’s potential as an ergogenic aid.

Prohormones and anabolic steroids

Testosterone and growth hormone are two primary hormones in the body that serve to promote gains in muscle mass (i.e., anabolism) and strength while decreasing muscle breakdown (catabolism) and fat mass [ 355 – 362 ]. Testosterone also promotes male sex characteristics (e.g., hair, deep voice, etc.) [ 356 ]. Low level anabolic steroids are often prescribed by physicians to prevent loss of muscle mass for people with various diseases and illnesses [ 363 – 374 ]. It is well known that athletes have experimented with large doses of anabolic steroids in an attempt to enhance training adaptations, increase muscle mass, and/or promote recovery during intense training [ 356 – 358 , 361 , 362 , 375 ]. Research has generally shown that use of anabolic steroids and growth hormone during training can promote gains in strength and muscle mass [ 355 , 360 , 362 , 368 , 371 , 376 – 383 ]. However, a number of potentially life threatening adverse effects of steroid abuse have been reported including liver and hormonal dysfunction, hyperlipidemia (high cholesterol), increased risk to cardiovascular disease, and behavioral changes (i.e., steroid rage) [ 378 , 384 – 388 ]. Some of the adverse effects associated with the use of these agents are irreversible, particularly in women [ 385 ]. For these reason, anabolic steroids have been banned by most sport organizations and should be avoided unless prescribed by a physician to treat an illness.

Prohormones (e.g., androstenedione, 4-androstenediol, 19-nor-4-androstenedione, 19-nor-4-androstenediol, 7-keto DHEA, and DHEA, etc.) are naturally derived precursors to testosterone or other anabolic steroids. Their use has been suggested to naturally boost levels of these anabolic hormones. While data is available demonstrating increases in testosterone [ 389 , 390 ], virtually no evidence exists demonstrating heightened training adaptations in younger men with normal hormone levels. In fact, most studies indicate that they do not affect testosterone and that some may actually increase estrogen levels and reduce HDL-cholesterol [ 378 , 389 , 391 – 396 ]. On a related note, studies have examined the ability of various ingredients to increase testosterone via inhibition of aromatase and 5-alpha-reductase [ 397 ]. Rohle et al. [ 398 ] and Willoughby et al. [ 399 ] reported that significant increases in free testosterone and dihydrotesterone occurred, but soft tissue composition either was not measured [ 398 ] or wasn’t changed as a result of supplementation [ 399 ]. Consequently, although there may be some potential applications for older individuals to replace diminishing androgen levels, it appears that prohormones have no training value. Since prohormones are “steroid-like compounds”, most athletic organizations have banned their use. Use of nutritional supplements containing prohormones will result in a positive drug test for anabolic steroids. Use of supplements (knowingly or unknowingly) containing prohormones have been believed to have contributed to a number of recent positive drug tests among athletes. Consequently, care should be taken to make sure that any supplement an athlete considers taking does not contain prohormone precursors particularly if their sport bans and tests for use of such compounds. Companies such as Informed Choice ( www.informed-choice.org ) and National Sanitation Foundation, NSF (aka, NSF Certified for Sport) www.nsf.org ) have developed assurance programs to test and screen various nutrition products. Moreover, several professional sporting organizations have incorporated language into the collective bargaining that requires all products provided by teams or sporting organizations must provide products that have achieved certain 3rd party approvals for safety, banned substances and/or label claims. It is noteworthy to mention that many prohormones are not lawful for sale in the USA since the passage of the Anabolic Steroid Control Act of 2004. The distinctive exception to this is dehydroepiandrosterone (DHEA), which has been the subject of numerous clinical studies in aging populations.

Sulfo-polysaccharides (myostatin inhibitors)

Myostatin or growth differentiation factor 8 (GDF-8) is a transforming growth factor known as a negative regulator of skeletal muscle hypertrophy [ 400 ]. In humans, inhibiting myostatin gene expression has been theorized as a way to prevent or slow down muscle wasting in various diseases, speed up recovery of injured muscles, and/or promote increases in muscle mass and strength in athletes [ 401 ]. Since 2010, no additional research has been published that examined the impact of any nutritional ingredient or strategy to inhibit myostatin expression. In humans, myostatin clearly plays a role in regulating skeletal muscle mass. For example, a study by Ivey and colleagues [ 401 ] reported that female athletes with a less common myostatin allele experienced greater gains in muscle mass during training and reduced atrophy during detraining. Interestingly, no such changes were reported for men. Willoughby and colleagues [ 402 ] supplemented untrained males with 1200 mg/day of Cytoseira Canariensis (a form of sea algae), a purported myostatin inhibitor, and reported no changes for fat-free mass, strength and blood concentrations of myostatin. These results were corroborated by Wilborn et al. [ 318 ] who reported no impact of sulfo-polysaccaride supplementation on body composition or performance changes. As it stands, there is currently no published data supporting the use of sulfo-polysaccharides or any other ingredient touted to act as a myostatin inhibitor for their ability to increase strength or muscle mass.

Tribulus terrestris

Tribulus terrestris (also known as puncture weed/vine or caltrops) is a plant extract that has been suggested to stimulate leutinizing hormone which stimulates the natural production of testosterone [ 403 ]. Consequently, tribulus is marketed as a supplement that can increase testosterone and promote greater gains in strength and muscle mass during training. In human research models, several studies have indicated that tribulus supplementation alone [ 404 , 405 ] or in combination with other segragotogues and androgen precusors [ 406 , 407 ] appears to have no effects on body composition or strength during resistance training.

Vanadyl sulfate (vanadium)

Vanadyl sulfate is a trace mineral that has been found to affect insulin-sensitivity (similar to chromium) and may affect protein and glucose metabolism [ 403 , 408 ]. In this regard, reports have highlighted the potential efficacy and support for vanadium to improve insulin sensitivity [ 409 ] and assist with the management of diabetes [ 410 ]. In relation to its potential ability to impact protein and glucose metabolism, vanadyl sulfate supplementation has been purported to positively impact strength and muscle mass [ 74 , 411 ]. However, no studies are available that support the ability of vanadyl sulfate supplementation to impact strength or muscle mass in non-diabetic individuals who are currently resistance training [ 412 , 413 ].

Zinc/magnesium aspartate (ZMA)

The main ingredients in ZMA formulations are zinc monomethionine aspartate, magnesium aspartate, and vitamin B-6. ZMA supplementation is based upon the rationale that zinc and magnesium deficiency may reduce the production of testosterone and insulin like growth factor (IGF-1). Consequently, ZMA supplementation is advocated for its ability to increase testosterone and IGF-1, which is further suggested to promote recovery, anabolism, and strength during training. Two studies with contrasting outcomes have examined the ability of acute ZMA administration to increase anabolic hormone concentrations. Initially, Brilla and Conte [ 414 ] reported that a zinc-magnesium formulation increased testosterone and IGF-1 (two anabolic hormones) leading to greater strength gains in football players participating in spring training while Koehler et al. [ 415 ] reported that ZMA supplementation increased serum zinc and excretion, but failed to change free and total testosterone levels. Wilborn et al. [ 416 ] had resistance trained males ingest a ZMA supplement or placebo in a double-blind fashion and resistance train for 8 weeks and found no change in free or total testosterone, strength or fat-free mass (via DXA). It is noted that previous deficiencies in zinc may negatively impact endogenous production of testosterone secondary to its role in androgen metabolism and steroid receptor interaction [ 417 ]. To this point, Brilla and Conte [ 414 ] did report depletions of both zinc and magnesium, thus increases in testosterone levels could have been attributed to deificient nutritional status rather than a pharmacologic effect. More research is needed to further evaluate the role of ZMA on body composition and strength during training before definitive conclusions can be drawn.

Performance enhancement supplements

Several nutritional supplements have been proposed to enhance exercise performance. Throughout this section, emphasis is placed upon results that directly measured some attribute of performance. In situations where a nutrient is purported to stimulate increases in fat-free mass and enhance performance (i.e., creatine), a large, more developed section is available while a shorter, more concise section is available in the other category. Table  3 categorizes the proposed ergogenic nutrients into: Strong Evidence to Support Efficacy and Apparently Safe, Limited or Mixed Evidence to Support Efficacy, Little to No Evidence to Support Efficacy and/or Safety.

ß-alanine

ß-alanine, a non-essential amino acid, has ergogenic potential based on its role in carnosine synthesis [ 12 ]. Carnosine is a dipeptide comprised of the amino acids, histidine and ß-alanine, that naturally occur in large amounts in skeletal muscles. Carnosine is believed to be one of the primary muscle-buffering substances available in skeletal muscle. Studies have demonstrated that taking four to 6 g of ß-alanine orally, in divided doses, over a 28-day period is effective in increasing carnosine levels [ 418 , 419 ], while more recent studies have demonstrated increased carnosine and efficacy up to 12 g per day [ 420 ]. According to the ISSN position statement, evaluating the existing body of ß-alanine research suggests improvements in exercise performance with more pronounced effects on activities lasting one to 4 min; improvements in neuromuscular fatigue, particularly in older subjects, and lastly; potential benefits in tactical personnel [ 12 ]. Other studies have shown that ß-alanine supplementation can increase the number of repetitions one can do [ 421 ], increase lean body mass [ 422 ], increase knee extension torque [ 423 ], and increase training volume [ 421 ]. In fact, one study also showed that adding ß-alanine to creatine improves performance over creatine alone [ 424 ]. While it appears that ß-alanine supplementation can improve performance, other studies have failed to demonstrate a performance benefit [ 425 , 426 ].

Caffeine is a naturally derived stimulant found in many nutritional supplements typically as guarana, bissey nut, or kola. Caffeine can also be found in coffee, tea, soft drinks, energy drinks, and chocolate. Caffeine has also been shown to be an effective ergogenic aid for aerobic and anaerobic exercise with a documented ability to increase energy expenditure and promote weight loss [ 14 ]. Research investigating the effects of caffeine on time trial performance in trained cyclists found that caffeine improved speed, peak power, and mean power [ 427 ]. Similar results were observed in a recent study that found cyclists who ingested a caffeine drink prior to a time trial demonstrated improvements in performance [ 428 , 429 ]. Studies indicate that ingestion of caffeine (e.g., 3–9 mg/kg taken 30–90 min before exercise) can spare carbohydrate use during exercise and thereby improve endurance exercise capacity [ 430 , 431 ]. In addition to the apparent positive effects on endurance performance, caffeine has also been shown to improve repeated sprint performance benefiting the anaerobic athlete [ 432 – 434 ]. Research examining caffeine’s ability to increase maximal strength and repetitions to fatigue are largely mixed in their outcomes. For example, Trexler, et al. [ 434 ] reported that caffeine can improve repeated sprint performance but failed to impact maximal strength and repetitions to fatigue using both upper-body and lower-body exercises. In agreement, Astorino and colleagues [ 435 ] revealed no change in upper-body and lower-body strength after resistance trained males ingested 6 mg/kg of caffeine. Similarly, Beck and investigators [ 436 ] provided resistance trained males with 201 mg caffeine (2.1–3.0 mg/kg) and reported no impact on lower-body strength, lower-body muscular endurance or upper-body muscular endurance. Maximal upper-body strength, however, was improved. In contrast, other studies have indicated that caffeine may favorably impact muscular performance. For example, Goldstein et al. [ 437 ] reported that caffeine ingestion (6 mg/kg) significantly increased bench press strength in a group of women but did not impact repetitions to fatigue. Studies by Duncan and colleagues [ 438 – 441 ] have examined the impact of caffeine on strength and endurance performance as well various parameters of mood state while performing maximal resistance exercise. Briefly, these authors have reported improvements in strength and repetitions to failure using the bench press [ 438 , 439 ] and other exercises [ 440 , 441 ]. In addition to potential ergogenic impact, these authors also reported that caffeine significantly improved various indicators of mood state [ 438 , 440 ], lowered ratings of perceived exertion and decreased perception of muscle pain [ 439 , 441 ] when acute doses of caffeine (5 mg/kg) were provided before maximal resistance exercise. As illustrated, when evaluating the research on caffeine for its ability to impact strength and muscular performance, the findings are equivocal, and, subsequently, more research is needed to better determine what situations may best predict caffeine’s ability to impact strength performance. For example, trained subjects have demonstrated more ergogenic effects compared to untrained subjects [ 442 , 443 ]. Also, people who drink caffeinated drinks regularly, however, appear to experience less ergogenic benefits from caffeine [ 444 ]. Some concern has been expressed that ingestion of caffeine prior to exercise may contribute to dehydration, although several studies have not supported this concern [ 430 , 445 , 446 ]. Caffeine, from anhydrous and coffee sources are both equally ergogenic [ 434 ]. Caffeine doses above 9 mg/kg can result in urinary caffeine levels that surpass the doping threshold for many sport organizations. In summary, consistent scientific evidence is available to indicate that caffeine operates as an ergogenic aid in several sporting situations.

One of the best ergogenic aids available for athletes and active individuals alike, is carbohydrate. Optimal carbohydrate in the diet on a daily basis, in the hours leading up to exercise, throughout exercise and in the hours after exercise can ensure endogenous glycogen stores are maintained and support many types of exercise performance [ 41 – 43 , 50 ]. In this respect, athletes and active individuals should consume a diet high in carbohydrate (e.g., 55–65% of calories or 5–8 g/kg/day) to maintain muscle and liver carbohydrate stores [ 41 , 50 , 54 ]. Research has clearly identified carbohydrate as an ergogenic aid that can prolong exercise [ 41 , 68 ]. For example, Below and colleagues [ 447 ] provided research that ingesting carbohydrate throughout a time to exhaustion protocol after nearly an hour of moderate intensity cycling can significantly extend the time cycling is performed. Moreover, Widrick et al. [ 129 ] systematically examined all four possible combinations of high and low pre-exercise intramuscular glycogen levels with and without carbohydrate provision before a standard bout of cycling exercise. When carbohydrate was provided, performance was improved. In addition to traditional endurance exercise models, Williams and Hawley [ 42 ] summarized the literature involving carbohydrate delivery and performance of team sports that are typically characterized by variable intensities and intermittent periods of heavy exertion and concluded that carbohydrate intake can increase performance. Pochmuller et al. [ 68 ] and Colombani et al. [ 69 ] have critically pointed to the duration of the involved exercise bout, the intensity of exercise involved, and the fasting status of the individuals as key factors that may impact exercise performance. Further, Burke and colleagues [ 23 , 50 ], Hawley et al. [ 43 ] and Rodriguez et al. [ 54 ] have all emphasized the importance of optimal carbohydrate delivery throughout various types of sport and recovery scenarios to support performance. Beyond ingestion, a growing body of literature has drawn attention to the potential impact of carbohydrate mouth rinsing as an ergogenic strategy. Initial work by Carter and colleagues [ 448 ] where they demonstrated an increase in time to exhaustion performance while cycling after rinsing (but not swallowing) the oral cavity with a carbohydrate solution versus a no carbohydrate rinse revealed that receptors in the brain might be linked to the mere presence of carbohydrate in the mouth, which subsequently can work to improve various types of exercise performance. While this concept is still emerging, some [ 449 – 452 ] but not all [ 453 – 455 ] of the studies have supported the ability of carbohydrate mouth rinsing to increase performance. Another carbohydrate manipulation strategy has included utilizing high molecular weight carbohydrates solutions, in contrast to traditional low molecular weight beverages, to theoretically accelerate glucose absorption and energy availability. Importantly, the majority of the literature suggests that utilizing a high molecular weight solution can impart changes in oxidized substrates, or patterns of fuel usage, but appears to have no ergogenic effect on performance in males or females [ 63 , 456 – 459 ].

As indicated earlier, creatine supplementation is a well-supported strategy to increase muscle mass and strength during training. However, creatine has also been reported to improve exercise capacity in a variety of settings [ 182 , 460 – 462 ]. Specifically, and as discussed by Kreider et al. [ 10 ], studies have documented improvements in: a) single and multiple sprints, b) work completed across multiple sets of maximal effort, c) anaerobic threshold, d) glycogen loading, e) work capacity, f) recovery, and g) greater training tolerance. Consequently, team sports, individual activities or sports that consist of high intensity, intermittent exercise such as soccer, tennis, basketball, lacrosse, field hockey and rugby can all benefit from creatine use [ 182 ]. Moreover, a 2009 study found that in addition to high intensity interval training creatine improved critical power [ 460 ]. Less research is available involving creatine supplementation and endurance exercise, but creatine’s ability to promote glycogen loading [ 463 ] and storage of carbohydrate [ 464 – 466 ], key fuels during endurance exercise, may translate into improved endurance exercise performance. Indeed, a 2003 study found that ingesting 20 g of creatine for 5 days improved endurance and anaerobic performance in elite rowers [ 467 ]. Since creatine has been reported to enhance interval sprint performance, creatine supplementation during training may improve training adaptations in endurance and anaerobic athletes, anaerobic capacity, and allow athletes to complete greater volumes of training at or above anaerobic threshold [ 468 , 469 ]. Notably, for athletes who struggle to maintain their body mass throughout their competitive season, creatine use may help athletes in this respect. Importantly and in addition to creatine being an effective ergogenic aid in a wide variety of sports, studies have documented these outcomes (improvements in acute exercise capacity, work completed during multiple sets and training adaptations) in adolescents [ 470 – 474 ], younger adults [ 231 , 424 , 462 , 475 – 483 ], and older individuals [ 484 – 490 ]. Regarding creatine and athletic performance, there appears to be a misunderstanding that creatine may result in muscle cramps and dehydration. However, based on many available studies, there is no clinical evidence that creatine supplementation will increase susceptibility of dehydration, muscle cramps, or heat related illness [ 196 , 491 ].

Sodium bicarbonate (baking soda)

During high intensity exercise, acid (H+) and carbon dioxide (CO 2 ) accumulate in the muscle and blood. The bicarbonate system is the primary means the body rids itself of the acidity and CO 2 via their conversion to bicarbonate prior to subsequent removal in the lungs. Bicarbonate loading (e.g., 0.3 g per kg taken 60–90 min prior to exercise or 5 g taken two times per day for 5 days) as sodium bicarbonate has been shown to be an effective way to buffer acidity during high intensity exercise lasting one to 3 min in duration [ 431 , 492 – 494 ]. Matson et al. [ 495 ] reported improvements in exercise capacity in events like the 400–800 m run while Lindh and colleagues [ 496 ] reported that bicarbonate can improve 200 m freestyle swimming performance in elite male swimmers. Similarly, studies have reported the ability of bicarbonate to improve 3 km cycling time trials [ 497 ]. Marriott et al. [ 498 ] published findings that sodium bicarbonate significantly improved intermittent running performance by 23% and reduced perceived exertion in male team-sport athletes. Interestingly, Percival and investigators [ 499 ] reported that sodium bicarbonate supplementation resulted in significantly higher levels of PGC-1-α, a key protein known to drive mitochondrial adaptations. Finally, a meta-analysis by Peart and investigators [ 500 ] involving sodium bicarbonate reported the overall treatment effect to be moderate at improving performance with nearly all measured ergogenic outcomes being influenced by the training status of the participants.

In addition, other studies have examined the potential additive benefit of ingesting sodium bicarbonate with either caffeine or beta-alanine. In this respect, Kilding et al. [ 497 ] reported significant independent effects of caffeine and bicarbonate on three-kilometer cycling time trial performance, but no additive benefit. Alternatively, Tobias and associates [ 501 ] also reported a significant improvement in upper-body power production in trained martial arts athletes after ingesting either beta-alanine or sodium bicarbonate, but noted a distinct synergistic improvement in upper-body power and performance when beta-alanine and sodium bicarbonate were ingested together. In contrast, Danaher et al. [ 502 ] had eight healthy males supplement with either beta-alanine, sodium bicarbonate or their combination for 6 weeks in a crossover fashion before completing a repeated sprint ability test while cycling. While buffering capacity was increased, performance was only improved when beta-alanine was provided. Due to the mixed outcomes and relative lack of available studies, more research is recommended examining the synergistic impact of sodium bicarbonate and other ingredients. It is important to highlight that a common complaint surrounding the ingestion of sodium bicarbonate is gastrointestinal distress, thus athletes should experiment with its use prior to performance to evaluate tolerance.

Sodium phosphate

Phosphate is best known as an essential mineral found in many common food sources (e.g., red meat, fish, dairy, cereal, etc.) with key functions in bone, cell membranes, RNA/DNA structure and as backbones of phosphocreatine and various nucleotides. In addition, phosphate has been suggested to operate in an ergogenic fashion due to its potential to improve oxygen transport through modulation of 2,3-diphosphoglycerate (DPG) and other lactic-acid-buffering components. Sodium phosphate (NaPO 4 ) supplementation has been reported in multiple studies to improve aerobic capacity by 5–12% [ 503 – 505 ], anaerobic threshold by 5–10% [ 504 – 507 ], mean power output [ 503 , 508 ] and intermittent running performance [ 509 – 511 ]. Collectively these studies have employed a dosing regimen that required 1 g of NaPO 4 to be taken four times daily for three to 6 days. Not all studies, however [ 512 – 514 ], have reported ergogenic outcomes while factors that impact phosphate absorption, training status and gender posed as potential reasons why supplementation has not universally impacted performance. Brewer and colleagues [ 513 ] reported modest (non-significant) effects of NaPO 4 supplementation on repeated supplementation regimens in trained cyclists completing a time trial. Furthermore, West and investigators [ 514 ] used a mixed gender cohort and concluded no change in VO 2 Max resulted after supplementation. Buck et al. [ 515 ] were the first to solely examine the impact of NaPO4 in female athletes when they had 13 trained female cyclists complete a 500-kJ time trial after supplementing with either 25, 50, or 75 mg/kg of NaPO 4 in a randomized, double-blind manner. No significant impact of supplementation was seen at any dosage leading the authors to conclude that females may not respond in the same manner as men. However, the same authors on two occasions [ 510 , 511 ] examined the impact of NaPO4 in female team sport athletes completing repeated bouts of sprint running and found that NaPO4 significantly improved best and total sprint times when compared to a placebo. Consequently, the impact of gender on the ergogenic potential of NaPO4 remains unclear with consistent benefits in females when repeated sprints are performed but no such benefits during time-trial work.

Water and sports drinks

Adopting strategies to limit the loss of body mass due to sweating is critical to maintain exercise performance (particularly in hot/humid environments). People engaged in intense exercise or work in the heat are commonly recommended to regularly ingest water or sports drinks (e.g., 12–16 fluid ounces every 10–15 min) with the overarching goal being to minimize the loss of body mass commonly seen as a result of exercising in a hot and humid environment [ 516 ]. Below and colleagues [ 447 ] demonstrated the independent ability of both fluid (no carbohydrate) and carbohydrate ingestion to significantly increase cycling performance. Moreover, when the two treatments were combined a synergistic impact on performance was observed. Studies show that ingestion of sports drinks during exercise in hot/humid environments can help prevent dehydration and improve endurance exercise capacity [ 517 – 519 ]. Of note and like carbohydrate, it appears that exercise factors such as the duration and intensity of the exercise bout operate as strong predictors of cycling time-trial performance [ 520 , 521 ]. Consequently, frequent ingestion of water and/or sports drinks during exercise is one of the easiest and most effective ergogenic aids due to its ability to support thermoregulation and reduce cardiovascular strain during prolonged bouts of exercise, particularly when completed in hot and humid conditions [ 162 , 516 ].

L-alanyl-L-glutamine

Operating under the same theoretical framework as glutamine, interest in supplementing with L-alanyl-L-glutamine has increased in recent years. The ingredient has two parts: L-alanine and L-glutamine, both of which are amino acids that are mainstays in the transamination processes involving amino acids. Rogero and colleagues [ 522 ] supplemented rats with L-alanyl-L-glutamine for the final 21 days of a six-week exercise training program. Supplementation did not impact time to exhaustion performance, but higher levels of glutamine were found when compared to a control group. Cruzat and Tirapequi [ 523 ] also reported increases in plasma and intramuscular glutamine along with an improved antioxidative profile in blood, muscle and liver tissue samples of laboratory rats. These results were extended in 2010 to also report an attenuation of inflammation and plasma creatine kinase levels in laboratory rats after exercise training [ 523 ].

Since 2010, five peer-reviewed studies have been published using human subjects. Hoffman and colleagues [ 524 ] reported, in a group of ten physically active males, that L-alanyl-L-glutamine increased time to exhaustion on a cycle ergometer when exposed to mild dehydration stress. Two years later, the same research group reported that rehydration with L-alanyl-L-glutamine after 2.3% dehydration in a basketball scrimmage led to an improvement in basketball skill performance and visual reaction time when compared to water [ 525 ]. A 2016 study indicated that L-alanyl-L-glutamine maintained reaction time in an upper and lower-body activities after an exhaustive bout of treadmill running [ 526 ]. Finally, a 2015 paper determined that L-alanyl-L-glutamine significantly improved treadmill running performance when compared to no hydration [ 527 ]. Collectively this research indicates that L-alanyl-L-glutamine at dosages ranging 300–1000 mg per 500 mL of fluid can favorably influence hydration status and performance when compared to no fluid ingestion or water only ingestion.

Arachidonic acid

Arachidonic acid (ARA) is a long-chain polyunsaturated fatty acid (20:4, n-6) that resides within the phospholipid bi-layer of cell membranes at concentrations that are dependent upon dietary intake [ 528 ]. ARA is not found in high amounts in the typical American diet [ 529 ]. However, as little as 1.5 g per day of supplementation over a 50-day period has been shown to increase tissue cell membrane stores of ARA [ 530 ]. In skeletal muscle, there is evidence that ARA drives some of the inflammatory response to strength training via enhanced prostaglandin signalling [ 531 ]. Specifically, exercise liberates ARA from the muscle cell membrane via phospholipase A 2 activation. Resultant free intracellular ARA is subsequently converted into certain prostaglandins (i.e., PGE 2 or PGF 2α ) via cyclooxygenase (COX) enzymes [ 532 ], and these prostaglandins can signal associated receptors in an autocrine and paracrine manner to up-regulate signalling associated with increases in muscle protein synthesis. Roberts and colleagues [ 533 ] were the first group to examine the impact of ARA supplementation on changes in strength and body composition. Over an eight-week period, resistance-trained, college-aged males were supplemented in a double-blind fashion with either a placebo or ARA at a dosage of 1 g per day in conjunction with 90 g/day of whey protein. A significant increase in anaerobic peak power was found in the ARA group, but no other changes in strength or body composition were found. The second study by DeSouza et al. [ 534 ] investigated the effects of ARA supplementation (0.6 g/d vs. placebo) in strength-trained college-aged males for 8 weeks with concomitant resistance training and without protein supplementation. These authors reported that lean body mass (2.9%, p  < 0.05), upper-body strength (8.7%, p  < 0.05), and anaerobic peak power (12.7%, p  < 0.05) significantly increased only in the ARA group. Mitchell and colleagues [ 535 ] have also published data in 19 resistance-trained men who supplemented, in a double-blind, placebo-controlled fashion, with 1.5 g per day of ARA for 4 weeks and found that ARA supplementation did not impact acute changes in muscle protein synthesis and other mechanistic links to protein translation. The authors concluded that ARA supplementation did not support a mechanistic link between ARA supplementation and short-term anabolism, but may increase translation capacity. Given the limited human data and inconsistent nature (two positive outcomes, one negative outcome) of the findings regarding the efficacy of ARA, it is too early to recommend ARA at this time. In this respect, more chronic human studies testing different doses of ARA supplementation are needed to fully examine its safety and potential efficacy as a performance enhancing or muscle building aid. From a safety perspective and due to ARA being a known pro-inflammatory fatty acid, use of ARA may be contraindicated in populations that have compromised inflammatory health (i.e., inflammatory bowel syndrome, Chron’s disease, etc.).

Ingestion of BCAA (e.g., 6–10 g per hour) with sports drinks during prolonged exercise has long been suggested to improve psychological perception of fatigue (i.e., central fatigue). Accordingly, Mikulski and investigators [ 536 ] used 11 endurance trained men to examine the impact of ingesting 16 g of BCAAs and 12 g of ornithine aspartate over a 90-min cycling exercise bout and found that the amino acid combination significantly improved reaction time, but no ergogenic impact was seen when BCAAs were ingested independently. Although a strong rationale and data exist to support an ergogenic outcome, mixed outcomes currently prevail as other studies have failed to report an ergogenic impact of BCAAs [ 247 , 537 ]. Consequently, more research is needed to fully determine the ergogenic impact, if any, of BCAAs. An important point to highlight surrounding BCAAs is the growing body of literature supporting their ability to mitigate outcomes surrounding muscle damage. In this respect, multiple studies have investigated and offered support for BCAA’s ability to promote recovery, mitigate soreness and attenuate losses in force production [ 249 , 250 , 537 , 538 ].

Citrulline (2-Amino-5-(carbamoylamino)pentanoic acid or L-Carnitine) is endogenously produced from ornithine and carbamoyl phosphate in the urea cycle. In the body, citrulline is efficiently recycled into arginine for subsequent nitric oxide production through the citrulline-nitric oxide cycle [ 539 ]. Unlike arginine, citrulline catabolism is limited in the intestines [ 540 ] as well as its extraction from hepatic tissue [ 541 ] resulting in the majority of citrulline passing into systemic circulation before conversion to arginine [ 542 ]. Due to this and its non-competitive uptake for cell transport [ 542 ], oral citrulline supplementation has been shown to be more effective in increasing arginine [ 543 , 544 ] and activation of nitric oxide synthase (NOS) [ 544 ] as well as various biomarkers of nitric oxide [ 545 ]. Multiple studies have employed aerobic exercise models to examine citrulline’s impact on performance. Suzuki et al. [ 546 ] showed that 2.4 g/day of L-citrulline for 7 days increased plasma nitric oxide metabolites, plasma arginine and 4-km time trial performance. Using a finger flexor exercise model and P31 nuclear magnetic resonance spectroscopy, Bailey and colleagues [ 547 ] reported that 7 days of citrulline (6 g/day) significantly increased plasma arginine and nitrite levels and significantly improved VO 2 kinetics and exercise performance. However, not all studies reported an ergogenic effect whereby Cunniffe et al. [ 548 ] reported no impact of 12 g of citrulline malate on the performance of a single bout of high-intensity cycling. In addition to aerobic exercise research, three studies examined the impact of an 8-g citrulline dose while resistance training on various performance outcomes [ 549 – 551 ]. One study [ 550 ] evaluated the effects on the number of repetitions performed for chin-ups, reverse chin-ups, and push-ups to failure in trained males. A second study [ 551 ] evaluated the effect of citrulline supplementation on the number of repetitions performed for five sequential sets (60% 1RM) to failure on the leg press, hack squat, and leg extension exercises in trained males. The third study [ 549 ] evaluated the effects of citrulline supplementation on the number of repetitions performed during six sets each of bench press and leg press exercises to failure at 80% 1RM in trained females. In all three studies, citrulline malate was shown to significantly increase performance during upper- and lower-body multiple-bout resistance exercise performance. Alternatively, Cultrufello and colleagues [ 552 ] reported that a 6 g dose of L-citrulline failed to impact both aerobic and anaerobic indicators of exercise performance. The role of malate in combination with citrulline is largely undetermined. Since malate is an important tricarboxylic acid cycle intermediate, this could possibly account for improvements in muscle function [ 553 ]. Therefore, it is presently unclear whether these benefits can be solely attributed to citrulline, as well as what role citrulline may play in aerobic and anaerobic performance.

Research exploring the impact of essential amino acids with various forms of exercise has exploded. To date, it is well accepted that ingestion of at least 2 g of the essential amino acid, leucine, is required to stimulate cellular mechanisms controlling muscle hypertrophy [ 225 , 554 ] and that ingestion of 6–12 g of a complete essential amino acid mixture are needed to maximize muscle protein synthesis [ 201 – 208 , 555 ]. However, their impact on performance remains largely unexplored. While sound theoretical rationale exists and multiple acute study designs provide supportive evidence, it is currently unclear whether following this strategy would lead to greater training adaptations and/or whether EAA supplementation would be better than simply ingesting carbohydrate and a quality protein following exercise. Moreover, very little research is available that has examined the ability of EAAs to impact exercise performance. For these reasons, many authors and review articles have encouraged the prioritization of intact protein sources over ingestion of free form amino acids [ 11 , 13 , 54 , 222 ] to promote accretion of fat-free mass, but, as mentioned, the impact of this recommendation on performance changes remains undetermined.

Ingesting glycerol with water has been reported to increase fluid retention, and maintain hydration status [ 556 – 558 ]. Theoretically, this should help athletes prevent dehydration and improve thermoregulatory and cardiovascular changes. Although studies indicate that glycerol can significantly enhance body fluid, results are mixed on whether it can improve exercise capacity [ 166 , 559 – 564 ]. Regarding endurance performance Coutts and investigators [ 565 ] had ten trained endurance athletes complete an Olympic distance triathlon under both placebo and glycerol hyperhydration (1.2 g/kg) + 25 mL/kg fluid solution) 2 h before completion of each triathlon and reported that completion time was significantly improved with glycerol hyperhydration over placebo. These findings were corroborated by Goulet et al. [ 566 ] when they had six endurance-trained subjects hyperhydrate with glycerol or water 2 h before a prolonged (2 h) bout of cycling at 65% VO 2 max in hot conditions (26-27 °C) followed two-minute intervals at 80% VO 2 max and concluded that glycerol hyperhydration significantly improved performance. In contrast, Marino et al. [ 567 ] reported that a similar glycerol hyperhydration protocol did not improve the total distance covered when moderately trained cyclists completed a variable-intensity cycling protocol. Additionally, Goulet et al. [ 568 ] combined a hyperhydration strategy (1.2 g/kg glycerol + 26 mL/kg water) 2 h before commencing a two-hour cycling bout at 66% VO 2 max and 25 °C with consuming (500 mL/hour) a sports drink and reported that glycerol hyperhydration failed to impact cardiovascular or thermoregulatory functions as well as endurance performance. McKenna and investigators [ 569 ] were one of the only research groups to examine glycerol’s potential to impact anaerobic power after glycerol hyperhydration. After following a double-blind hyperhydration protocol, male collegiate wrestlers lost 3% of their body mass from fluid and completed an anaerobic test where no impact on performance was found. Variable outcomes surrounding glycerol continue to undermine its potential and the ability to offer a recommendation for its use. Consequently, as pointed out by Goulet et al. [ 556 ], it is concluded that more research needs to be completed to work through the nuance surrounding glycerol’s potential efficacy, a key point previously summarized by Nelson et al. [ 570 ].

For several years, beta-hydroxy-beta-methyl-butyrate (HMB) has received interest for its ability to enhance training adaptations and performance while also delaying or preventing muscle damage [ 15 , 168 , 571 ]. Initial work by Nissen and colleagues [ 171 ] showed significant increases in lean body mass and strength with doses 1.5 and 3 g/day in untrained males, with the 3 g dose showing additional benefits over the lower dose. Gallagher and colleagues [ 169 ] indicated that a dose of 38 mg/kg/day (approximately 3 g/day) promoted improvements in fat-free mass, peak isometric force and isokinetic torque production, while no changes in maximal strength were seen. In agreement, Thomson and researchers [ 572 ] had 22 resistance trained men supplement, in a double-blind fashion, with either HMB or placebo for 9 weeks and concluded that HMB was responsible for a significant increase in lower-body strength. Not all studies, however, have provided support. For example, Kreider et al. [ 175 ] used a dose-response, placebo-controlled approach and concluded that three or 6 g of calcium-HMB did not impact body composition or strength adaptations in individuals experienced with resistance exercise after 4 weeks of supplementation and resistance training. Similarly, Hoffman and colleagues [ 573 ] reported that HMB supplementation failed to improve anaerobic power production in collegiate football players, a conclusion which aligns with other previous studies [ 172 , 574 ]. Differences in training regimens (intensities), randomization, and supervision varied in the initial studies and may have contributed to the mixed results. HMB appears to have the greatest effects on performance when training intensity is maximized.

While many of the previous studies have examined, with mixed results, the ergogenic potential of calcium-HMB supplementation in active, recreationally active individuals, Durkalec-Michalski and colleagues completed three investigations [ 178 – 180 ] that all sought to determine the impact of calcium-HMB supplementation in different athlete types. For instance, HMB supplementation (3 g/day) in elite rowers over a 12-week period significantly improved aerobic (VO 2 max, time to reach ventilatory threshold) performance markers and decreased fat mass when compared to changes seen with placebo [ 178 ]. Later, Durkalec-Michalski and Jeszka [ 179 ] required 58 highly trained males to supplement with calcium-HMB (3 g/day) for 12 weeks. In this study, fat-free mass increased and fat mass decreased along with multiple markers of aerobic capacity when HMB was provided in comparison to a placebo. Most recently, HMB supplementation over 12 weeks in highly-trained combat sport athletes significantly increased (in comparison to placebo) several indicators of aerobic and anaerobic exercise performance [ 180 ]. The recent studies by Durkalec-Michalski and colleagues confirmed earlier works by Vukovich [ 575 ] and Lamboley [ 576 ] that HMB does have a positive effect on increasing aerobic capacity.

HMB is available as calcium-HMB and as free acid. In comparison to calcium HMB, HMB-free acid shows greater and faster absorption (approx. 30 min vs. 2–3 h) [ 577 ]. Much of the initial research used calcium-HMB with largely mixed outcomes while studies using the free acid form are more limited. Studies by Wilson and colleagues using the free acid form have indicated robust changes in strength, vertical jump power and skeletal muscle hypertrophy while heavy resistance training alone [ 578 ] and in combination with supplemental ATP [ 579 ], but others have critically questioned these outcomes [ 580 ]. A recent systematic review by Silva and investigators [ 581 ] concluded that the free acid form of HMB may improve muscle and strength and attenuate muscle damage when combined with heavy resistance training but stated that more research is needed before definitive conclusions can be determined.

Nitrate supplementation has received much attention due to their effects on vasodilation, blood pressure, improved work efficiency, modulation of force production, and reduced phosphocreatine degradation [ 582 – 584 ] all of which can potentially improve sports performance. Nitrate supplementation is most commonly consumed two to 3 h prior to exercise as beetroot juice or sodium nitrate [ 585 ] and is prescribed in both absolute and relative amounts ranging from 300 to 600 mg [ 585 ] or 0.1 mmol per kilogram of body mass per day, respectively [ 583 , 586 ]. These dosing amounts appear to be well tolerated when consumed as both supplemental [ 587 ] and supplemental sources [ 588 ] without significant alterations in hemodynamics or clinical boundaries of hepatorenal and muscle enzyme status [ 478 , 589 ]. Supplementing highly trained cyclists with sodium nitrate (10 mg per kilogram of body mass) significantly reduced VO 2 peak without influencing time to exhaustion or maximal power outputs [ 590 ]. Additionally, 600 mg of nitrate supplementation (given 2 h prior) non-significantly improved the performance of a 500-m time trial performance in elite-level kayak athletes by 2 s [ 591 ]. Of practical significance, it should be noted that first place and last place in the 2008 Beijing Olympics, was separated by 1.47 s in the 500-m men’s canoe/kayak flatwater race. Amateur cyclists at simulated altitude (~ 2500 m) observed improved 16.1 km time trial performance with a concomitant decrease in oxygen consumption after beetroot juice (310 mg nitrate) supplementation [ 592 ]. Not all findings, however, have reported performance benefits with nitrate supplementation. Nitrate supplementation (~ 385 mg nitrate) 2.5 h before a 50-mile time trial in well-trained cyclists failed to improve performance [ 593 ], which was also reported by MacLeod et al. [ 594 ] after examining nitrate supplementation (~ 400 mg nitrate) on 10-km time trial performance in normoxia or simulated altitude (~ 2500 m). In well-trained runners, nitrate supplementation (~ 430 mg nitrate) did not improve performance during an incremental exercise test to exhaustion (simulated altitude 4000 m) or a 10-km time trial (simulated altitude, 2500 m) [ 595 ] and Nyakayiru et al. [ 596 ] reported no impact of nitrate supplementation on changes in VO 2 and time trial performance in highly trained cyclists. Other studies have also reported an additive or synergistic effects of high-intensity intermittent exercise, endurance exercise, or resistance training when nitrate supplementation is combined with sodium phosphate [ 511 ], caffeine [ 597 ], or creatine [ 478 ], respectively. It is important to mention that dietary nitrates have a health benefit in some, but not all populations [ 598 ]. Daily consumption of beetroot juice (~ 320–640 mg nitrate/d) significantly decreased resting systolic blood pressure in older adults by approximately 6 mmHg [ 599 , 600 ]. Nitrate supplementation (560 mg – 700 mg nitrate) significantly increased blood flow to working muscle and exercise time in older adults with peripheral artery disease [ 601 ] as well as significantly improved endothelial function via increased flow-mediated dilation and blood flow velocity in older adults with risk factors of cardiovascular disease [ 602 ]. Collectively, these results indicate that nitrate supplementation may improve aerobic exercise performance and cardiovascular health in some populations.

Post-exercise carbohydrate and protein

Ingesting carbohydrate with protein following exercise has been a popular strategy to heighten adaptations seen as part of a resistance training program. The rationale behind this strategy centers upon providing an energy source to stimulate MPS via key signal transduction pathways. Additionally, carbohydrate intake will impact insulin status which could promote MPS, limit protein breakdown or both [ 603 – 605 ]. Furthermore, combining carbohydrate with protein can heighten glycogen resynthesis rates, particularly when carbohydrate intake is not optimal [ 120 ] and can improve muscle damage responses after exhaustive exercise [ 606 ]. A key point for readers to consider when interpreting findings from this literature is the amount of protein, essential amino acids or leucine being delivered by the protein source [ 11 ]. In the last few years many studies have agreed that post workout supplementation is vital to recovery and training adaptations [ 133 , 230 , 232 , 607 , 608 ]. However, the need for adding carbohydrate to protein to maximize hypertrophic adaptations continues to be questioned. For example, Staples and investigators [ 605 ] used an acute study design involving stable isotope methodology to investigate the impact of adding 50 g of carbohydrate to 25 g of whey protein ingestion after a single bout of lower body resistance exercise. The authors concluded that the combination of carbohydrate and protein was no more effective at stimulating muscle protein synthesis or blunting rates of muscle protein breakdown than protein alone. Furthermore, Hulmi and colleagues [ 609 ] had participants resistance train for 12 weeks and supplement with equivalent doses of whey protein, carbohydrate or whey protein + carbohydrate while having strength and body composition assessed. Overall, changes in strength were similar in all groups while changes in fat-free mass were greater in the protein group when compared to the carbohydrate group. Fat mass was found to significantly decrease in both groups that contained protein in comparison to carbohydrate, but no differences between the two protein-containing groups were noted. In conclusion, these findings underscore the importance of ingesting adequate protein to stimulate resistance training adaptations. Whether or not the addition of carbohydrate can heighten these changes at the current time seems unlikely. This outcome, however, should not distract the reader from appreciating the fact that optimal carbohydrate delivery will absolutely support glycogen recovery, aid in mitigating soreness and inflammation and fuel other recovery demands.

Quercetin is a flavonoid commonly found in fruits, vegetables and flowers, and is known for having some health benefits with therapeutic use. In addition, quercetin has been purported in both animal and human models to improve endurance performance. In this respect, Cureton and colleagues [ 610 ] supplemented 30 recreationally active, but not highly trained men in a double-blind fashion to ingest either quercetin (1 g/day) or placebo. No changes in total work performed, substrate utilization, or perception of effort were found after supplementation. Similarly, Bigelman and investigators [ 611 ] supplemented ROTC cadets with either 1 g of quercetin or a placebo and concluded that VO 2 max was unchanged as a result. These results correspond with the outcomes of other studies that failed to document ergogenic potential for quercetin [ 612 , 613 ]. In contrast, Nieman et al. [ 614 ] supplemented untrained adult males with 1 g of quercetin in a double-blind fashion for 2 weeks and reported that treadmill performance and markers of mitochondrial biogenesis were improved. Similarly, Patrizio et al. [ 615 ] used a resistance exercise model and reported quercetin may improve neuromuscular performance while Davis et al. [ 616 ] had 12 study participants supplement with either quercetin and placebo and found that quercetin may improve VO 2 max and endurance capacity. A meta-analysis was completed by Pelletier and researchers [ 617 ] to summarize the potential impact of quercetin supplementation on endurance performance. This analysis involved seven published studies representing 288 research participants. Only in untrained participants was quercetin found to significantly increase endurance performance. A 2011 meta-analysis by Kressler et al. [ 618 ] drew a similar conclusion whereby they indicated quercetin does have benefit, but the size of this effect is trivial and small. Consequently, more research needs to be completed to better identify what situations may exist that support quercetin’s ability to impact exercise performance.

Taurine is an amino acid found in high abundance in human skeletal muscle [ 619 , 620 ] derived from cysteine metabolism that plays a role in a wide variety of physiological functions [ 621 – 623 ]. Studies have indicated that training status (higher in trained vs. untrained muscle, reviewed in [ 624 ]) and fiber type (higher in type I vs. type II, reviewed in [ 619 ]) impact the amount of taurine found in muscle. It has been reported in some [ 625 , 626 ] but not all studies [ 627 , 628 ] that taurine may improve exercise performance and mitigate recovery from damaging and stressful exercise [ 629 ]. In recent years, many studies have examined the impact of taurine ingestion on various types of exercise performance. In accordance with previous work, ergogenic outcomes related to taurine administration continue to be mixed. Milioni and investigators [ 628 ] failed to show an improvement in performance with a 6 g dose of taurine while completing high-intensity treadmill running. Similarly, Balshaw et al. [ 625 ] indicated that taurine failed to positively impact 3-km running performance in trained runners. In contrast, a 2017 study by Warnock et al. [ 630 ] reported that a 50 mg/kg dose of taurine outperformed caffeine, placebo and caffeine + taurine on performance changes after repeated Wingate anaerobic capacity tests. Finally, a 2018 meta-analysis by Waldron et al. [ 631 ] reported that single daily dosages ranging from one to 6 g for up to 2 weeks can significantly improve endurance exercise performance in a range of study participants. Two studies [ 632 , 633 ] have been completed that examined taurine’s ability to mitigate decrements associated with muscle damage and resistance exercise performance. Notably, oral ingestion at a dosage of 50 mg/kg for 14 days prior to damage and for 7 days after damage significantly increased strength, and decreased soreness and markers of muscle damage [ 633 ]. Finally, studies have also supported the ability of taurine to function in an anti-oxidative role, which may promote an improved cellular environment to tolerate exercise stress [ 634 , 635 ]. While more research continues to be published involving taurine, the consensus of these outcomes continue to be mixed regarding taurine’s potential to enhance physical performance.

Arginine is known as a conditionally essential amino acid which has been linked with the ability to increase exercise performance, increase growth hormone production, support immune function, increase training tolerance and promote accretion of fat-free mass [ 273 , 636 ]. Several studies have sought to examine the ergogenic potential of arginine using both endurance and resistance exercise models with largely mixed results. For example, Greer and investigators [ 637 ] examined the ability of arginine + alpha-ketoglutarate and reported that the combination did not significantly impact muscle endurance and significantly reduced the number of chin-ups completed. Similar outcomes were found by Aguiar et al. [ 638 ] in older women whereby arginine supplementation failed to impact muscle performance. Sunderland et al. [ 639 ] supplemented 18 endurance-trained cyclists for 28 days with either arginine (12 g/day) or corn starch and concluded that arginine did not impact VO 2 Max or ventilatory threshold. In accordance, several other studies have failed to positively report on the ability of arginine to operate as an ergogenic aid [ 278 , 640 – 644 ]. Alternatively, a few studies have provided evidence of ergogenic potential for arginine. For example, Campbell and researchers [ 271 ] supplemented 35 resistance-trained males in a double-blind fashion with arginine (2 g) + alpha-ketoglutarate (2 g) or a placebo and concluded that maximal upper-body strength and wingate peak power were significantly increased after supplementation. Similarly, Bailey and colleagues [ 645 ] concluded that acute arginine supplementation reduced the oxygen cost of moderate-intensity exercise and increased tolerance to high-intensity training. Moreover, Pahlavani et al. [ 646 ] supplemented male athletes with arginine in a double-blind fashion and concluded that arginine supplementation significantly increased sport performance. As it stands, most of the published literature that has examined the ability of arginine to operate in an ergogenic fashion has failed to report positive outcomes. While more research is certainly indicated, consumers should exercise caution when using arginine to enhance exercise performance.

Carnitine is produced endogenously by the liver and kidneys and plays a pivotal role in lipid metabolism. Consequently, many are led to believe that carnitine ingestion will increase the concentration of endogenous carnitine, thereby increasing lipid metabolism and decrease adipose reserves. To date, the majority of the data continues to suggest that carnitine supplementation does not markedly affect muscle carnitine content [ 647 – 649 ], fat metabolism [ 648 , 650 , 651 ], exercise performance [ 648 , 649 , 652 , 653 ], or weight loss in overweight [ 650 , 654 ], obese [ 651 , 655 , 656 ] or trained subjects [ 657 ]. For example, Burrus and investigators [ 658 ] had ten cyclists ingest combinations of carbohydrate and carnitine while completing a 40-min ride at 65% VO 2 peak before completing an exhaustion ride at 85% VO 2 Peak. No differences in power outputs or times to exhaustion were found with cycling at 85% VO 2 Peak. Of note, studies have suggested that co-ingesting carnitine with carbohydrate can lead to significant increases in intramuscular carnitine [ 659 , 660 ]. Later, Wall and colleagues [ 661 ] reported that endurance exercise performance was improved and improvements in fuel selection appeared to occur. While interesting, more research is needed regarding changes in performance before further recommendations can be made.

As outlined above, a strong theoretical framework exists for glutamine’s ability to help an individual tolerate stress, particularly when relying on animal studies. A close examination into the available human research on glutamine makes it more challenging to characterize glutamine’s potential. Theoretically, glutamine supplementation during training should enhance gains in strength and muscle mass, but evidence in this respect has not been consistent. Glutamine supplementation has been shown to improve glycogen stores which could go on to impact certain types of exercise performance [ 331 ] and two recent studies suggest that glutamine provision may help support recovery from damaging resistance exercise. In this respect, Street and colleagues [ 662 ] concluded that adding glutamine (0.3 g/kg) to a carbohydrate drink significantly improved muscle soreness and force production, but did not impact changes in creatine kinase, when compared to carbohydrate only ingestion. A similar outcome was found by Legault and colleagues [ 337 ] who reported that glutamine supplementation significantly lowered perceived soreness levels and led to improved recovery of force production after a damaging bout of eccentric muscle contractions. From an ergogenic perspective, limited research is available, but Antonio et al. [ 334 ] reported that 0.3 g/kg glutamine ingestion did not impact the number of repetitions completed with the leg press or bench press exercises. Consequently, minimal research is available to support glutamine’s ability to operate as an ergogenic aid.

Inosine is a building block for DNA and RNA that is found in muscle. Inosine possesses important roles that may enhance training and/or exercise performance [ 663 ]. Although there is some theoretical rationale, available studies indicate that inosine supplementation has no apparent effect on aerobic or anaerobic exercise performance [ 664 – 666 ].

Medium chain triglycerides

Medium chain triglycerides (MCT’s) are shorter chain fatty acids known to readily enter the mitochondria and be converted to energy through beta-oxidation [ 667 ]. Studies are mixed as to whether MCT’s are ergogenic and can serve as an effective source of fat during exercise [ 667 – 671 ]. A 2001 study found that 60 g/day of MCT oil for 2 weeks did improve running performance [ 672 ]. Additionally, Van Zyl and colleagues [ 671 ] reported that while MCTs negatively influenced cycling time trial performance when ingested alone in comparison to carbohydrate ingestion, performance was improved when MCTs were combined with carbohydrate. Using a similar exercise model, Goedecke et al. [ 668 ] also reported that MCT administration throughout a two-hour moderate intensity cycle ride resulted in a similar performance in completing a 40-km time trial by trained cyclists. A similar outcome was also reported by Vistisen et al. [ 673 ]. Beyond equivocal findings, Goedecke et al. [ 674 ] and Jeukendrup et al. [ 667 ] both reported ergolytic outcomes of MCT administration on sprint performance in trained cyclists and cycling time trial performance, respectively, while the incidence of gastrointestinal complaints increased in both studies. These findings have been confirmed by others that MCT oils are not sufficient to induce positive training adaptations and may cause gastric distress [ 675 , 676 ]. Consequently, it does not appear likely that MCT favorably impacts acute exercise performance and no evidence exists that training adaptations may be positively impacted either, while multiple studies have reported that MCT ingestion may cause gastrointestinal upset and decrease exercise performance.

Ribose is a 5-carbon carbohydrate that is involved in the synthesis of adenosine triphosphate (ATP) and other adenine nucleotides. Clinical studies have shown that ribose supplementation can increase exercise capacity in heart patients [ 677 – 681 ] leading to the development of theories that it can operate as ergogenic aid for athletes. Of the available research, most fail to show an ergogenic value for ribose supplementation on exercise capacity in healthy untrained or trained populations [ 682 – 684 ]. A 2006 study [ 685 ] investigated the effects of supplementing with either ribose or dextrose over 8 weeks on rowing performance and concluded that ribose was outperformed by the dextrose control [ 685 ]. Kreider and associates [ 684 ] and Kerksick and colleagues [ 686 ] investigated ribose supplementation on measures of anaerobic capacity in trained cyclists and concluded ribose had no positive impact on performance. In 2017, Seifert and investigators [ 687 ] had 26 healthy subjects supplement with either 10 g of ribose or 10 g of dextrose for 5 days while completing a single bout of interval exercise and a two-minute power output test. When splitting the participants into high vs. low oxygen uptake levels, the people with low peak VO 2 experienced significant increases in mean and peak power output along with reductions in ratings of perceived exertion and creatine kinase. No such changes were reported in individuals with high peak VO 2 . As it stands, clinical findings provide support while studies in healthy, trained populations generally fail to report a positive outcome for ribose supplementation. Healthy individuals with lower fitness levels may afford some benefit.

Supplements to promote general health

In addition to the supplements previously described, several nutrients have been suggested to help athletes stay healthy during intense training. For example, the American Medical Association has recommended that all Americans ingest a daily low-dose multivitamin in order to ensure that people get a sufficient level of vitamins and minerals in their diet [ 688 , 689 ]. Although daily vitamin and mineral supplementation has not been found to improve exercise capacity in athletes, it may make sense to take a daily vitamin supplement for health reasons. Vitamin D is often recommended to athletes, especially those participating in indoor sports or in cloudy geographies [ 690 ]. Although direct evidence linking vitamin D with performance is equivocal, it is clear that vitamin D has a role in regulating immune function, cardiovascular health, and growth and repair. Dosing should be dependent upon baseline levels, which can be measured by any physician [ 691 ]. Glucosamine and chondroitin have been reported to slow cartilage degeneration and reduce the degree of joint pain in active individuals which may help athletes postpone and/or prevent joint problems [ 692 , 693 ]. Meanwhile, other ingredients including undenatured type II collagen (UC-II) may be helpful as well although more research is needed involving athletic applications [ 694 , 695 ]. Supplemental vitamin C, glutamine, echinacea, quercetin, and zinc have been reported to enhance immune function [ 125 , 696 – 699 ]. However, consuming carbohydrate during prolonged strenuous exercise attenuates rises in stress hormones and appears to limit the degree of exercise-induced immune depression [ 699 ]. Similarly, although additional research is necessary, vitamin E, vitamin C, selenium, alpha-lipoic acid and other antioxidants may help restore overwhelmed antioxidant defenses exhibited by athletes [ 700 ]. One countering argument against higher doses is the potential for these to interfere with adaptive responses to training [ 699 ]. Finally, the omega-3 fatty acids docosahexaenoic acid (DHA) and eicosapantaenoic acid (EPA), in supplemental form, are now endorsed by the American Heart Association for heart health in certain individuals stemming from initial scientific statements made in 2002 [ 701 ]. This supportive supplement position stems from: 1) an inability to consume cardio-protective amounts by diet alone; and, 2) the mercury contamination sometimes present in whole-food sources of DHA and EPA found in fatty fish. For general health, dosing recommendations range from 3000 mg-5000 mg daily of deep, cold water fish [ 702 ]. Consequently, prudent use of these types of nutrients at various times during training may help athletes stay healthy and/or tolerate training to a greater degree.

High intensity exercise can compromise an athlete’s immune health. Infection risk and exercise workload follow a J-Shape curve with moderate intensity exercise reducing the infection risk, and high intensity exercise actually increasing the risk of infection [ 703 ]. Immune suppression in athletes further worsens by the psychological stress, foreign travel, disturbed sleep, environmental extremes, exposure to large crowds or an increase exposure to pathogens due to elevated breathing during exercise or competition. Athletes have several nutritional options to reduce the risk and symptoms of upper respiratory tract infections, including probiotics and baker’s yeast beta-glucan. Beta-glucan is a natural gluco polysaccharide derived from the cell walls of highly purified yeast ( Saccharomyces cerevisiae ) and has been shown to significantly decrease upper-respiratory tract infection symptoms in men and women participating in the Carlsbad marathon [ 704 ]. Probiotics, often referred to as “friendly” or “good” bacteria, are live microorganisms which when administered in adequate amounts confer a health benefit on the host. An estimated 70% of our immune system is located in your digestive system indicating the importance of a balanced gut microflora on immune health. Probiotics have been shown to reduce the number, duration and severity of upper-respiratory tract infections and gastrointestinal distress in the general population and in athletes, certain strains of probiotics have been shown to significantly reduce the number of upper-respiratory tract infection episodes a well as their severity [ 705 ]. Health benefits of probiotics are strain specific and dose dependent, and some strains have failed to show beneficial effects in athletes [ 706 ]. Also, consuming carbohydrate during prolonged strenuous exercise attenuates rises in stress hormones and appears to limit the degree of exercise-induced immune depression [ 699 ].

Several factors operate as cornerstones to enhance athletic performance and optimize training adaptations including the consumption of a balanced, nutrient and energy dense diet, prudent training, and obtaining adequate rest. Use of a limited number of nutritional supplements that research has supported to improve energy availability (e.g., sports drinks, carbohydrate, creatine, caffeine, β-alanine, etc.) and/or promote recovery (carbohydrate, protein, essential amino acids, etc.) can provide additional benefit in certain instances. Dietitians and sport nutritionists should stay up to date on current research regarding the role of nutrition on exercise so they can provide honest and accurate information to their students, clients, and/or athletes about the role of nutrition and dietary supplements on performance and training. Furthermore, these professionals should actively participate in exercise nutrition research, write unbiased scholarly reviews for journals and lay publications, and help disseminate the latest research findings to the public. Through these actions, consumers and other professionals can make informed decisions about appropriate methods of exercise, dieting, and/or whether various nutritional supplements can affect health, performance, and/or training. In all situations, individuals are expected and ethically obligated to disclose any commercial or financial conflicts of interest during such promulgations. Finally, companies selling nutritional supplements or promoting exercise, diet or supplementation protocols should develop scientifically based products, conduct research on their products, and honestly market the results of studies so consumers can make informed decisions.

Acknowledgements

Preparation of these documents require the help of many individuals. For starters, the authors would like to acknowledge all of the authors from the 2010 publication: Lem Taylor, Bill Campbell, Anthony Almada, Conrad Earnest, Doug Kalman, Brian Leutholtz, Hector Lopez, Ron Mendel, Marie Spano, Darryn Willoughby, Tim Ziegenfuss and Joey Antonio. In particular, the authors are indebted to the assistance provided by Jonathan Manfre, Esq who assisted with critical edits to many parts related to regulation and to Darryn Willoughby who offered valuable additions to sections of manuscript. The authors would like to thank all the research participants, graduate students, and researchers that contributed to the body of research cited in this comprehensive review. Finally, we would like to thank the late Mel Williams who inspired many of the authors to pursue research evaluating the role of nutrition on exercise and performance. Any discussion provided for any given supplement is not intended to be nor should it be construed as any form of endorsement by any of the authors, the ISSN or the respective universities, corporations or entities that each author is affiliated. Individuals interested in trying some of these nutritional recommendations should do so only after consulting with their personal physician.

Abbreviations

Authors’ contributions.

RBK and CMK contributed most of the content and served as senior editors on the paper; CDW and MDR provided substantial support in revising and drafting several sections. RC, SMK, MBC, JND, EG, MG, RJ, LML, ASR, and RW all contributed to the development of selected parts of the manuscript. All authors reviewed and approved the final draft of this manuscript. RBK is the corresponding author.

This paper was reviewed by the International Society of Sports Nutrition Research Committee and represents the official position of the Society.

Not applicable.

Authors of this paper have not received any financial remuneration for preparing or reviewing this paper. All authors report the following competing interests: CMK consults with and receives external funding from companies who sell supplemental protein, has received remuneration from companies for delivering scientific presentations at conferences and writes online, print and other media on topics related to exercise, nutrition and protein for related companies. Has served as an expert witness and provided testimonies related to exercise, supplementation and nutrition. CDW has received external funding from supplement companies to do research, served on multiple advisory boards for supplement companies, and has served as a consultant, advisor, and spokesperson for various nutrition companies. MDR has received academic and industry funding related to dietary supplements, served as a non-paid consultant for industry and received honoraria for speaking about topics discussed in this paper. He currently has no patents, stock or ownership in any companies doing business on topics discussed in this paper. RC is the attorney for numerous companies in the dietary supplement industry and has received payment for consultancy and the writing of lay articles discussing nutritional supplements. SMK has served as a paid consultant for industry; has received honoraria for speaking at conferences and writing lay articles about topics discussed in this paper; receives royalties from the sale of several exercise and nutrition related books; and, receives commission and has stock in companies that sell products produced from several ingredients discussed in this paper.

JND has no conflicts of interest to report. MBC has received academic and industry funding related to dietary supplements, served as a consultant for industry and received honoraria for speaking about topics discussed in this paper. EG has no conflicts of interest to report. MG has received external funding and nutritional product from companies who sell protein supplements and has received remuneration from companies for presenting scientific based nutritional supplement and exercise research at professional conferences. RJ has received grants to evaluate the efficacy and safety of dietary and food ingredients, serves on scientific advisory boards, and has served as an expert witness, legal and scientific consultant. LML has received academic and industry funding related to dietary supplements and honoraria from speaking engagements on the topic and has received payment for consultancy and the writing of lay articles discussing nutritional supplements. ASR has received grants to evaluate the efficacy of dietary supplements, serves as a scientific advisor for sports nutrition companies, and received remuneration from companies for presenting evidenced-based nutritional supplement and exercise research at professional conferences. RW has received industry funds for consultancy and employment related to dietary supplement development and marketing and currently works as the Chief Science Officer for Dymatize Nutrition. RBK has received externally-funded grants from industry to conduct research on dietary supplements and has served as a scientific and legal consultant.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Chad M. Kerksick, Email: ude.doownednil@kciskrekc .

Colin D. Wilborn, Email: ude.bhmu@nrobliwc .

Michael D. Roberts, Email: ude.nrubua@4200rdm .

Abbie Smith-Ryan, Email: ude.cnu.liame@htimsbba .

Susan M. Kleiner, Email: moc.renielksrd@nasus .

Ralf Jäger, Email: [email protected] .

Rick Collins, Email: moc.qsebmgc@snillocr .

Mathew Cooke, Email: [email protected] .

Jaci N. Davis, Email: ude.bhmu@sivadj .

Elfego Galvan, Email: [email protected] .

Mike Greenwood, Email: ude.umat@62doowneergm .

Lonnie M. Lowery, Email: ude.noinutnuom@mlyrewol .

Robert Wildman, Email: moc.ezitamyd@namdliwr .

Jose Antonio, Email: ude.avon@938aj .

Richard B. Kreider, Email: ude.umat@redierkbr .

IMAGES

  1. Health And Nutrition Articles 2018

    sports nutrition research paper

  2. Amazon.com: Sports Nutrition: A Handbook for Professionals, Sixth

    sports nutrition research paper

  3. FREE 10+ Nutrition Assessment Forms in PDF

    sports nutrition research paper

  4. (PDF) Sports Nutrition for Young Athletes

    sports nutrition research paper

  5. 📗 Paper Example on Sports Nutrition

    sports nutrition research paper

  6. Sports Nutrition Essay Example

    sports nutrition research paper

VIDEO

  1. hs nutrition question paper 2024 solve ।। class 12 nutrition question paper 2024 answers

  2. sports nutrition 0/100 #nutrition #atheletics #diet

  3. Sports Nutrition: Functional Food Advantages

  4. SPORTS NUTRITION: FROM SCIENCE TO RECOMMENDATIONS SPONSORED BY GSSI: HYDRATION, Pahnke, M

  5. TOUGHEST PAPER Reaction 😪😭 I MATHS CLASS 12th PAPER I CBSE PAPER

  6. Applied Nutrition 2nd semester 2022 solve past paper //Part=1 #kmu #bsn #nutrition #pastpaper

COMMENTS

  1. Sports Nutrition: Diets, Selection Factors, Recommendations

    Among athletes, nutrition plays an important role since the regimen and composition of the diet are associated with success in sports [23,24].Concerns about weight and body shape strongly influence food choices for the general population [] and have a similar effect on athletes, where attempts to achieve their goals are associated with external data on physique, weight, and performance [].

  2. Journal of the International Society of Sports Nutrition

    The purpose of this study was to determine... Lee M. Margolis, Marques A. Wilson, Claire C. Whitney, Christopher T. Carrigan, Nancy E. Murphy, Adrienne Hatch-McChesney and Stefan M. Pasiakos. Journal of the International Society of Sports Nutrition 2021 18 :56. Research article Published on: 10 July 2021. Full Text.

  3. (PDF) Sports Nutrition and Performance

    Abstract and Figures. Nutrition plays an essential role on sports performance. Following an adequate nutrition pattern determines winning the gold medal or failing in the attempt. That is why it ...

  4. ISSN exercise & sports nutrition review update: research

    14 Exercise & Sports Nutrition Lab, Human Clinical Research Facility, Texas A&M University, College Station, TX, USA. [email protected]. ... This paper is an ongoing update of the sports nutrition review article originally published as the lead paper to launch the Journal of the International Society of Sports Nutrition in 2004 and updated in ...

  5. Current and Novel Reviews in Sports Nutrition

    A major emphasis in all of the papers was a focus on strengths and weaknesses for various sports nutrition strategies, and insights for future research. Kerksick et al. [ 11 ] defined the role that proper doses of plant proteins can have in supporting health, the environment, and exercise training adaptations.

  6. New Opportunities to Advance the Field of Sports Nutrition

    Sports nutrition is a relatively new discipline; with ~100 published papers/year in the 1990s to ~3,500+ papers/year today. Historically, sports nutrition research was primarily initiated by university-based exercise physiologists who developed new methodologies that could be impacted by nutrition interventions (e.g., carbohydrate/fat oxidation by whole body calorimetry and muscle glycogen by ...

  7. International society of sports nutrition position stand: nutrient

    The International Society of Sports Nutrition (ISSN) published the first position stand devoted to the practice of nutrient timing in 2008 [ 1 ]. Consequently, this paper has been accessed approximately 122,000 times. In the past nine years, multiple lines of research have explored questions directly related to the timing of nutrients that ...

  8. Full article: ISSN exercise & sports nutrition review update: research

    Sports nutrition is a constantly evolving field with hundreds of research papers published annually. In the year 2017 alone, 2082 articles were published under the key words 'sport nutrition'. ... originally called the Sports Nutrition Review Journal). This paper provides a definition of ergogenic aids and dietary supplements and discusses ...

  9. Performance Nutrition for Athletes

    The first paper [] addresses the intriguing topic of translating sports performance nutrition research into the real world and ultimately the chances of it helping an athlete maximize their performance to reach the "podium."The authors present a framework they call the "Paper-2-Podium Matrix" which provides several criteria to critically evaluate performance nutrition-related research ...

  10. PDF ISSN exercise & sports nutrition review update: research & recommendations

    nutrition, the ISSN Exercise & Sports Nutrition Review: Research & Recommendations has been up-dated. The initial version of this paper was the first publication used to help launch the Journal of the International Society of Sports Nutrition (JISSN, ori-ginally called the Sports Nutrition Review Journal). This paper provides a definition of ...

  11. Sports Nutrition for Optimal Athletic Performance and Health ...

    The paper on nutrition and athlete bone health [] also stresses the need for more athlete-specific research, especially as it relates to longer-term bone health (e.g., risk of osteopenia and osteoporosis) and shorter-term risk of bony injuries.Bone is a nutritionally modified tissue and generally benefits from weight-bearing activities, although not all athletes engage in weight-bearing sports.

  12. ISSN exercise & sports nutrition review update: research

    Sports nutrition is a constantly evolving field with hundreds of research papers published annually. In the year 2017 alone, 2082 articles were published under the key words 'sport nutrition'. Consequently, staying current with the relevant literature is often difficult. This paper is an ongoing update of the sports nutrition review article originally published as the lead paper to launch ...

  13. Free Full-Text

    A major emphasis in all of the papers was a focus on strengths and weaknesses for various sports nutrition strategies, and insights for future research. Kerksick et al. [ 11 ] defined the role that proper doses of plant proteins can have in supporting health, the environment, and exercise training adaptations.

  14. Sports Nutrition for Optimal Athletic Performance and Health: Old, New

    The paper on nutrition and health at altitude has been written by several scientists renowned for their work examining strategies to enhance adaptation, improve performance and maintain health in athletes living and training at low-to-moderate altitudes (1600-2400 m). Much of the existing altitude research was conducted at high to extreme ...

  15. Full article: ISSN exercise & sport nutrition review: research

    Athletes and active individuals should consume a diet high in carbohydrate (e.g., 55 - 65% of calories or 5-8 grams/kg/day) in order to maintain muscle and liver carbohydrate stores [ 1, 3 ]. Research has clearly identified carbohydrate is an ergogenic aid that can prolong exercise [ 3 ].

  16. Sports Nutrition

    Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications. ... Sports, diet, nutrition ...

  17. From Paper to Podium: Quantifying the Translational Potential of

    Abstract. Sport nutrition is one of the fastest growing and evolving disciplines of sport and exercise science, demonstrated by a 4-fold increase in the number of research papers between 2012 and 2018. Indeed, the scope of contemporary nutrition-related research could range from discovery of novel nutrient-sensitive cell-signalling pathways to ...

  18. International Society of Sports Nutrition Position Stand: protein and

    Position statement The International Society of Sports Nutrition (ISSN) provides an objective and critical review related to the intake of protein for healthy, exercising individuals. Based on the current available literature, the position of the Society is as follows: 1) An acute exercise stimulus, particularly resistance exercise, and protein ingestion both stimulate muscle protein synthesis ...

  19. Structure and trends of international sport nutrition research between

    Sport nutrition as a research topic has attracted great attention in the scientific literature in the field of sport and exercise science. Various systematic reviews or meta-analyses have been conducted on numerous aspects of sport nutrition because of its complex nature that gains interest worldwide [1-6]. Apart from the classical and ...

  20. Performance Nutrition for Athletes

    The first paper [ 1] addresses the intriguing topic of translating sports performance nutrition research into the real world and ultimately the chances of it helping an athlete maximize their performance to reach the "podium.". The authors present a framework they call the "Paper-2-Podium Matrix" which provides several criteria to ...

  21. International society of sports nutrition position stand: nutritional

    1. Methods. The International Society of Sports Nutrition (ISSN) position stands are invited papers on topics the Journal of the ISSN (JISSN) Editors and Research Committee identifies as being of interest to JISSN readers.

  22. ISSN exercise & sport nutrition review: research & recommendations

    Sports nutrition is a constantly evolving field with hundreds of research papers published annually. For this reason, keeping up to date with the literature is often difficult. This paper is a five year update of the sports nutrition review article published as the lead paper to launch the JISSN in 2004 and presents a well-referenced overview of the current state of the science related to how ...

  23. ISSN exercise & sports nutrition review update: research

    Sports nutrition is a constantly evolving field with hundreds of research papers published annually. In the year 2017 alone, 2082 articles were published under the key words 'sport nutrition'. Consequently, staying current with the relevant literature is often difficult.