Skip to main content
  • Systematic Review
  • Open access
  • Published:

The Relationship Between Undulatory Underwater Kick Performance Determinants and Underwater Velocity in Competitive Swimmers: A Systematic Review

Abstract

Background

Undulatory underwater swimming (UUS) has become an integral component of the start and turn phases in competitive swimming allowing higher velocities than can be achieved swimming at the surface. An understanding of the most important determinants for UUS performance and how these can be optimised to different swimmers is poorly understood.

Objective

The aim of this systematic review was to systematically assess the current peer-reviewed literature on the relationship between UUS performance determinants and underwater velocity in competitive swimmers.

Methods

An electronic search using AusSportMed, Embase, PubMed, SPORTDiscus and Biomechanics and Medicine in Swimming was performed. The methodological quality of the studies was evaluated using a biomechanics-specific checklist developed by Hindle and colleagues (Sports Med Open. 5(1):49, 2019. 10.1186/s40798-019-0222-z).

Results

Twenty-five studies met the eligibility criteria. While UUS velocity was nearly perfectly related (r > 0.90) to foot resultant acceleration and kick frequency, several other biomechanical factors were also significant correlates. UUS velocity and frequency were typically higher in high-performance swimmers and during prone versus dorsal positions. UUS velocity, kick frequency and kick amplitude were also significantly correlated with high angular velocities of the hip, knee and ankle joints and knee range of motion.

Conclusion

While there appears to be evidence supporting some performance variables to be related to UUS, future research should examine how to optimise the kinematic and kinetic characteristics with respect to the imposed task constraints and organism constraints between swimmers. Additional research should also investigate the effect of biomechanically informed interventions to improve UUS performance.

Registration

Open Science Framework.

Key Points

  • A range of kinematic and kinanthropometric parameters are strongly correlated with undulatory underwater velocity; however, foot resultant acceleration, kick frequency, kick amplitude, vertical toe velocity and knee angular velocity appear to be the greatest predictors of high UUS velocity.

  • Swimmers should perform the glide at approximately 0.4 m underwater at all velocities above 1.9 ms-1 to gain maximum drag reduction benefits, where a 15-18 % reduction in underwater total drag was found when compared to swimming at the surface.

  • An athlete’s optimal movement combination when performing undulatory underwater swimming may be different to others, as each swimmer needs to exploit their own organism constraints to maximise propulsive impulse while simultaneously reducing drag in response to the task and environmental constraints.

Introduction/Background

Pool swimming is a foundation Olympic sport incorporating four recognised strokes: freestyle, butterfly, backstroke and breaststroke, with events ranging from 50 to 1500 m [2, 3]. Competitive pool swimming events can be divided into four distinct phases; the start, free swimming, turn and finish. The phases can be defined as: the start (time to 15 m), free swimming (parts of the race not including start, turn or finish), turn (5 m into the wall and 10 m out) and finish (5 m into the wall on the last lap). With early swimming research focused primarily on the free swimming component, more recent studies have recognised the contribution of the other phases of the race as important determinants for overall performance [4, 5]. Undulatory underwater swimming (UUS) has become an integral component of the start and turn phases since the 1980 Olympic Games in Moscow, where swimmers began to prolong the underwater phase, applying undulatory swimming to minimise the loss of velocity until the initiation of the above water stroke [6, 7]. The timing and types of UUS are also key factors for minimising deceleration from initial velocity from the dive or wall start [8]. Specifically, the undulatory kick is preferable to flutter kicking as it reduces deceleration before the initiation of the free swimming portion [8]. International Swimming Federation (FINA) regulations currently allow a swimmer to remain underwater for 15 m during the start and turn phases of the race [9], meaning the UUS may contribute up to 30% of race distance during a standard long course event [10]. Performance determinants of the UUS can be broadly categorised as ways to minimise resistance drag and/or to improve propulsive force production.

The total drag forces experienced by the swimmer reflect wave, skin and profile drag [11]. Because the density of water is approximately 800 times higher than that of air, the drag force in water is also higher [8]. The UUS promotes improved propulsive efficiency compared to swimming at the surface by eliminating wave drag [12, 13]. As swimming velocity, frontal surface area and frontal shape are the primary determinants of form drag force, the influence of frontal area and shape can be reduced by adopting a streamlined, horizontal position, with both arms overstretched and held together in front of the head while performing UUS [14,15,16]

Human underwater undulatory motion is dependent on the swimmer’s ability to produce propulsive forces, which are primarily generated by a wave running cephalocaudally along the athlete’s body [17]. The undulatory wave increases in amplitude at each subsequent body segment in a whip-like action, where momentum is transferred from larger body segments to smaller ones [10, 18, 19] and to the water resulting in a propulsive impulse [20]. Underwater velocity is equal to product of kick frequency and horizontal distance per kick, and the optimal interaction between these two factors [16, 21]. The cyclic vertical motion of the lower limbs caused by the body wave has the greatest amplitude at the toes, which is related to vortex creation and thrust production [10, 15, 19]. It has been stated that body wave velocity and vertical toe velocity could potentially explain differences in UUS performance, with implications for improving athlete technique [15].

An athlete’s ability to optimise propulsion and minimise resistance in the UUS may also be influenced by their anthropometry, range of motion and flexibility. The tendency for swimmers to be tall and lean assists in reducing drag, with their long limbs contributing to greater stroke length and/or kick amplitude [22]. However, there are some anthropometric differences between male and female swimmers and those who specialise in different strokes and distances [22]. A swimmer’s anthropometry may also have conflicting effects on their movement efficiency. When performing the UUS, the increase in displacement from the cranial to caudal body segments is not smooth due to the limited number of rotational joints available in the human body, compared to aquatic animals [23]. Range of motion and flexibility may then play an important role in creating and maintaining the optimal body position during UUS by allowing the swimmer to reduce resistance forces and/or increase propulsive force production [17].

Despite the frequent use of the UUS in training and competition, the factors determining its effectiveness in high-performance swimmers remain somewhat unclear [24]. This relative lack of understanding may partially reflect the most recent review on UUS published by Connaboy et al. [25] that only included six studies and the major changes that have occurred in the dive start since that time. Specifically, high-performance swimmers now use a kick or track start technique on the OSB11 start block (OMEGA, Zurich, Switzerland), which was first introduced in 2010. It provides a number of biomechanical and performance advantages for swimmers during the starting phase [26,27,28]. The purpose of this systematic review was to systematically assess the current peer-reviewed literature on the relationship between UUS kinematic, kinetic and kinanthropometric (anthropometric and physical fitness) factors that may influence UUS velocity in swimmers.

Methods

Experimental Approach to the Problem

A review protocol for this paper was developed using the Preferred Reporting Items for Systematic Reviews and Meta-Analyses’ (PRISMA) guidelines on reporting items for a systematic review and the associated PRISMA checklist [29]. The protocol was registered with Open Science Framework (https://osf.io/), and a set of inclusion and exclusion criteria were developed prior to undertaking the search process, as summarised in Table 1.

Table 1 Inclusion and exclusion criteria for the systematic review

Search Strategy

A structured literature search, using established search terms appropriate for each of the following databases: PubMed, SPORTDiscus, Embase and AusSportMed, was initially carried out with assistance from the University Faculty librarian on the 23 January 2020 and repeated on the 4 May 2021. A search of the database International Symposium on Biomechanics and Medicine in Swimming was then completed on the 6 December 2021. The search strategy included Medical Subject Headings (MeSH) terms and key words related to the primary concepts of the overall research question, using the PICO (population, intervention, comparison/control and outcome) approach. [The full search used included: swim* AND (dolphin OR undulat* OR underwater) AND kick* as the basis for searches in the other databases. For PubMed: (swim* OR "Swimming"[Mesh]) AND (dolphin OR undulat* OR underwater) AND kick*; for SPORTDiscus: (swim* OR DE "SWIMMING" OR DE "INDIVIDUAL medley" OR DE "LONG distance swimming" OR DE "MEDLEY relay (Swimming)" OR DE "MIDDLE distance swimming" OR DE "SWIMMING competitions" OR DE "SWIMMING for people with disabilities" OR DE "SWIMMING for women") AND (dolphin OR undulat* OR underwater) AND kick*; for Embase: (swim* OR 'swimming'/exp) AND (dolphin OR undulat* OR underwater) AND kick*; and for AusSportMed: swim* AND (dolphin OR undulat* OR underwater) AND kick*].

Risk of Bias and Quality Assessment

A risk of bias and quality assessment was undertaken by two independent reviewers. As no standard checklist appeared to be entirely suitable for eligible cross-sectional biomechanical studies in this review, a checklist developed by Hindle and colleagues [1] and subsequently used by several other authors was utilised [1, 30, 31]. When any disagreements in the scoring between reviewers occurred, a consensus meeting was held to establish an agreement. An item was scored as one if the article provided sufficient evidence in support of the criteria, and zero where the criteria were not met. A total risk of bias score was calculated for each article and categorised using the methods of Davids and Roman [32], with articles scoring ≥ 67% considered as having a low risk of bias, articles scoring in the range of 34–66% considered as having a satisfactory risk of bias and articles scoring ≤ 33% considered as having a high risk of bias.

Data Extraction and Analysis

Database results were combined, and duplicates were deleted. Titles and abstracts were screened for key words, and citations were then categorised into inclusion and exclusion sets to determine total number of records for synthesis. Data were extracted before risk of bias assessment was conducted using Endnote as the reference management software package. Dual screening was then carried out, using all inclusion and exclusion criteria by two independent reviewers, by titles and abstracts. For articles that appeared to meet the inclusion criteria, or where it was not clear, full-text reports were examined. The same two reviewers then independently screened the full text against the eligibility criteria, and any discrepancies were discussed with an experienced third researcher to reach a consensus. Reasons for any exclusions were recorded.

Data extracted from each eligible study included descriptive information of the study population, including number, age, performance level (regional club, state, national and international) and stroke identified. Data regarding the study design, UUS parameters measured, outcome(s) measured, statistical relationships and main findings reported were also extracted. For studies that reported correlations between the reported outcomes, qualitative descriptions of strength of the correlations were provided based on Hopkins using the following criteria: < 0.1, trivial; 0.1–0.3, small; 0.3–0.5, moderate; 0.5–0.7, large; 0.7–0.9, very large; > 0.9, nearly perfect [4].

Results

Study Characteristics, Methodology and Review Statistics

Figure 1 represents the article review process based on the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) guidelines [29]. The literature search identified 338 studies. Of the 46 articles retained for full-text screening, 22 articles were identified as being adherent to the inclusion criteria and three additional studies were identified through other sources, yielding a total of 25 studies included in the systematic review. The results of the search process are illustrated in a PRISMA flow chart in Fig. 1.

Fig. 1
figure 1

PRISMA flow chart [29]

Risk of Bias and Quality Assessment

All studies clearly stated the objectives or purpose of the study, described the testing methods and results and had a study design that adequately tested the hypothesis. The majority of studies clearly described the characteristics of the population and provided sufficient information that would allow the reader to make an unbiased assessment. All articles were classified as having a satisfactory or low risk of bias (Table 2) [1].

Table 2 Method for assessing quality and risk of bias

Study Results and Data Synthesis

Table 3 summarises the demographic characteristics of the participants as well as the primary methodological characteristics of the study, number of trials and the biomechanical analysis used in each study. The sample sizes of the included articles ranged from 1 to 40 participants, with mean ages between 16 and 26 years. Studies used various methodologies to test their hypotheses, with 1 to 3 trials of 10–20 m maximum UUS performance efforts the most common approach assessed [6, 7, 10, 14,15,16,17, 24, 33,34,35,36,37,38,39,40,41,42,43,44]. Nineteen of the 25 included studies utilised 2D kinematic analyses [6, 7, 10, 14,15,16,17, 24, 33, 34, 37, 39,40,41,42,43, 45,46,47] with 8 of these studies collecting data of how active kinanthropometric measures such as joint range of motion may be related to UUS performance [6, 10, 15, 17, 33,34,35, 41]. Of the 6 studies that performed a kinetic analysis, 4 studies calculated net and drag force data and reported the coefficient of drag [42,43,44,45], and 2 used a swimming flume to examine flow characteristics during UUS [38, 46].

Table 3 Summary of study participants, demographic characteristics, trials performed, kinematic, kinetic and kinanthropometric analysis

Correlation and regression analyses were one of the most commonly performed statistical approaches used in the eligible studies (Table 4). Across the nine studies that performed correlation analyses, UUS velocity was found to be significantly correlated with at least one kinematic outcome in eight [6, 7, 10, 15, 34, 36, 37, 41] and at least one kinanthropometric outcome in four studies [6, 10, 24, 41]. The most common significantly correlated kinematic variables to UUS velocity were kick frequency [6, 7, 37], vertical toe velocity [10, 15], knee angular velocity [15, 34] kick amplitude [36, 37] and angle of attack for the trunk, thigh and foot [36, 37]. While the strength of these correlations was typically strong to nearly perfect in magnitude, there were between-study variation for kick frequency (r = 0.43–0.90). Two of these studies also performed regression analyses to obtain greater insight into determinants of UUS velocity [34, 36].

Table 4 Summary of study kinematic and kinanthropometric correlations to underwater kick velocity

Two of the three studies reporting kinanthropometric results utilised active range of motion measures during the UUS [6, 10], whereas another study assessed single joint muscular strength measures [24]. The strength of the correlations for the active range of motion measures (r = 0.45–0.90) was larger than those for the muscular strength measures (r = 0.47–0.53), strong to very large, and moderate to strong, respectively.

Table 5 provides a summary of some of the primary biomechanical descriptors of the UUS phase, with an emphasis on how these may change as a function of body position, horizontal distance from the starting block and across different levels of swimmers. The majority of studies reported UUS velocity [6, 7, 16, 17, 24, 33,34,35,36,37, 39,40,41, 48, 49], kick frequency [6, 7, 16, 24, 33,34,35,36,37,38, 40, 41, 49] and kick amplitude [6, 7, 15, 24, 34,35,36,37, 40, 41, 48, 49].

Table 5 Summary of study underwater kick performance variables data most frequently correlated with velocity

Alves et al. [6], and Arellano et al. [33] collected data in prone, dorsal and ventral body positions and compared UUS velocity, kick frequency, kick amplitude and knee range of motion across the different body positions. The UUS velocity (prone: 1.46 and 1.69 m s−1; dorsal: 1.42 and 1.67 m s−1) and frequency (prone: 2.35 and 2.22 Hz; dorsal: 2.30 and 2.25 Hz) were similar in both the prone and dorsal body positions, with both positions substantially greater than in the lateral position. Alves et al. [6] reported values for kick amplitude for each body position (prone: 0.50 m; dorsal: 0.55 m).

de Jesus et al. [35], Houel, Elipot, Andree and Hellard [36], and Houel et al. [37] sought to examine how UUS velocity, kick frequency and amplitude may change as a function of horizontal distance from the starting block. In backstroke starts, UUS velocity decreased from the first four kick cycles to the last four cycles prior to resurfacing at 15 m [35]. The reductions in UUS velocity appear to reflect declines in both kick amplitude and kick frequency that were observed from the first four to the last four kick cycles [35]. Houel and colleagues [36, 37] also saw a consistent decrease in velocity as the swimmer approached the 15 m mark.

Several studies also reported data for different levels of swimmers. As expected, the UUS velocity was typically greatest for international swimmers, followed by national and regional swimmers [17, 38, 39]. Higher UUS velocities were generally associated with higher kick frequencies and consistent kick amplitudes [7, 17, 24, 34, 38, 40]. Two studies examined other kinematic factors that may be related to UUS velocity, kick frequency or kick amplitude. Wang and Lui [39] reported that the international swimmers had a significantly greater UUS knee angular velocity than the regional-level swimmers; however, no numerical data were provided for either group of swimmers. Yamakawa et al. [49] also provided data for knee angular velocity during the up and down kick phases, as well as knee range of motion during the UUS in national-level male swimmers (Table 5).

The study conducted by Shimojo et al. [47] had 15 national swimmers (10 M, 5F) who perform two tasks, one that required 10 × 10 m UUS trials at different kick frequencies determined by a target sound, with the second task being identical with the exception of having no sound. The results, that were reported as timing error (s) and displacement error (%), indicated that providing a target sound as a form of auditory augmented feedback improved their timing (which may improve their ability to maintain an optimal kick frequency), but that this feedback had less influence on their displacement measures [47].

Four studies reported data in relation to the net, drag and/or reaction forces, as well as the coefficient of drag associated with aspects of the UUS (Table 6) [42, 43, 45]. Lyttle et al. [43] examined how the net and drag forces would change at different velocities when the swimmer was passively towed underwater at a depth of 0.6 m. As the velocity increased from 1.6 m s−1 up to 3.1 m s−1, the net force became increasingly negative as a result of the drag force increase in magnitude. While Jensen and McIIwain [45] reported a drag force and reaction force, the description of their methods was poor, with no other relevant data provided. Hochstein et al. [44] reported the coefficient of drag during the glide and undulatory phases of UUS and compared the respective results between computational fluid dynamics (CFD) and experimental conditions. The study saw flow field similarities; however, the experimental results for the undulatory phases were not specified. CFD showed a much larger coefficient of drag during undulatory underwater swimming compared to gliding [44].

Table 6 Summary of study kinetic data

Five of the 25 studies included in this systematic review reported results related to UUS hydrodynamic mechanisms, including vortices, jet flow and wake [10, 14, 38, 44, 46] (see Table 7). Two of these studies used a swimming flume [38, 46], whereas the other two studies obtained 2D kinematic data from 3 × 15 m trials to calculate the hydrodynamic outcomes [10, 14]. Hochstein et al. [44] compared hydrodynamic mechanisms between CFD and experimental conditions during the glide and undulatory phases.

Table 7 Summary of study hydrodynamics data

Discussion and Implications

The purpose of this systematic review was to critically appraise the current peer-reviewed literature on how biomechanical factors might influence UUS performance. The primary finding of the current systematic review was that a range of variables may be strong correlates or predictors of UUS velocity. A number of these UUS variables were also found to vary as a function of body position (prone and dorsal), horizontal distance from the starting block and level of swimmer, with such differences potentially explained as a result of variations in the vortices in the wake surrounding the swimmer. While this review provides a systematic analysis of our current understanding of the kinematics, kinetics and hydrodynamics impacting UUS, no long-term training intervention research exists to provide an understanding of how chronic changes in biomechanical and/or kinanthropometric factors may interact with each other and contribute to alterations in UUS velocity and overall swim performance.

Performance Determinants Correlated with UUS Velocity

As a cyclical human activity, it could be hypothesised that a kick frequency would result in greater UUS velocity. Consistent with this hypothesis, the results of this review indicated that a range of UUS kinematic and kinanthropometric outputs that either directly or indirectly correspond to kick amplitude and kick frequency characteristics had large (r = 0.5–0.7), very large (r = 0.7–0.9) or nearly perfect (r > 0.9) correlations to UUS velocity. The strongest correlations to UUS velocity were foot resultant acceleration (r = 0.94) and kick frequency (r = 0.90), reported by Alves et al. [6] in a group of six junior national swimmers. Very large correlations to UUS velocity were also found between several kinematic variables, peak vertical toe velocity (r = 0.85), shoulder angle (r = 0.80), body wave velocity (r = 0.78) and peak hip angular velocity (r = 0.73), and measures of the maximum knee flexion and extension angle (r = 0.88) and maximum knee extension angle (r = 0.84) [10, 15, 41].

However, the study conducted by Ikeda et al. [41] reports that there was no significant correlation between kick frequency and velocity (r = 0.28); however, a significant relationship between kick frequency and the duration of the deceleration phase (r = − 0.842) was observed. Such results suggested that the more time a swimmer spends in the up kick phase, the lower the kick frequency, which may be indicative of utilising a higher kick amplitude to drive thrust and propulsion. Evidently, the time of the deceleration phase correlated with the relative vertical velocity of the shoulder to greater trochanter at maximal swimming velocity.

Vertical Toe Velocity and UUS

Peak vertical toe velocity during the up kick and down kick can explain approximately 72.3% of the variance in UUS velocity (r = 0.85) [15]. Peaks in horizontal UUS velocity have been found to occur at the same time, or immediately after peaks in vertical toe velocity, and this increase in forward speed is observed to be more apparent during the down kick, illustrated by a strong correlation between foot resultant acceleration and UUS velocity (r = 0.94) [6, 10, 49]. The similar timing of the peak in vertical toe velocity and UUS velocity may reflect a strong relationship between the magnitude and timing of maximum vertical toe velocity, and that of horizontal acceleration [23, 33, 46, 50, 51].

Atkison et al. [10] reported in a study of regional and international male swimmers that average vertical toe velocity during the up kick showed the highest correlation to the swimmer’s centre of mass velocity (r = 0.63), suggesting that faster swimmers are more proficient at the up kick than slower swimmers [10]. Additionally, significant correlations were identified between up kick velocity and horizontal velocity (r = 0.99) as well as down kick velocity and horizontal velocity (r = 0.98) [10]. Higgs and colleagues [15] also found that vertical toe velocity (r = 0.85) and up kick duration (− 0.79), supported by results of Ikeda et al. [41] were significantly correlated with UUS velocity, as well as body wave velocity, peak hip extension angular velocity and mean knee flexion angular velocity [15, 41].

Angular Velocities and UUS

The up kick is characterised by hip extension and knee flexion [10], with better swimmers typically extending at the hip before initiating knee flexion [50] suggesting that a high hip angular velocity may have a positive influence on thrust production and performance [15]. Mean hip (DK: − 85.0 ± 33.2 deg s−1; UK: 79.0 ± 32.5 deg s−1) and peak hip (DK: − 191.9 ± 61.4 deg s−1; UK: 248.0 ± 44.8 deg s−1) angular velocity and mean knee (DK: 260.0 ± 28.9 deg s−1; UK: − 190.1 ± 43.6 deg s−1) and peak knee (DK: 532.8 ± 36.9 deg s−1; UK: − 237.5 ± 40.5 deg s−1) angular velocity values suggest that the knee action contributes more than the hip action during the down kick, whereas the contribution of the hip and knee is similar during the up kick [15]. These findings align with that of Arellano et al. [50] and Atkison et al. [10], where elite swimmers utilise hip extension in the up kick to a greater extent than novice performers, highlighting the importance of hip extension. Further, Ikeda et al. [41] identified in their study of nine elite male swimmers that faster swimmers moved their lower trunk with greater angular displacement during the acceleration phase (down kick) than the slower swimmers (fast swimmers: 25.2°, slow swimmers: 4.5°), where the angular displacement of the lower trunk was also correlated with the angular displacements of the shoulder, knee, upper leg and lower leg. Despite the importance of the lower trunk angular displacement, excessive lower trunk movement may increase frontal projection area which in turn creates greater water resistance [14, 41].

Connaboy et al. [34] found in their parsimonious model for maximal swimming velocity that maximum knee angular velocity, ankle angular velocity and knee range of motion accounted for a very large amount of variance in UUS velocity (adjusted r2 = 0.93). After removing the participant as the fixed factor from the analysis, the strength of the prediction was reduced (adjusted r2 = 0.40). As a result, maximum ankle angular velocity and knee range of motion were no longer statistically significant, leaving only maximum knee angular velocity as a significant predictor of UUS velocity. The results of these two regression analyses support the importance of maximum knee angular velocity during UUS, but also that an individual’s kinanthropometry and/or technique may influence the strength of the relationship between joint angular velocity and range of motion to that of their UUS velocity [34]. For example, a high mean knee angular velocity allows the swimmer to complete the latter phase of the up kick phase quickly to minimise their hydrodynamic resistance [15, 34] and also increase kick frequency by reducing the duration of each kick cycle. Subject-specific analysis [52] should be employed to consider the importance of individual UUS techniques when interpreting data [34].

Body Wave Velocity and UUS

A stepwise regression analysis revealed that maximum vertical toe velocity and body wave velocity were strong predictors of UUS velocity (r2 = 0.72), with an additional 5.2% explained by the mean body wave velocity (partial correlation r = 0.46) [15]. Undulatory locomotion is accomplished via oscillations that pass along the length of a swimming body, and an understanding of the temporal sequencing between these oscillations as they occur is fundamental to appreciating how UUS velocity can be optimised. The oscillations generate a wave that transfers momentum to the surrounding fluid resulting in propulsive impulse and thrust [15, 20]. The composition of the body oscillations and their relationships determine the shape and velocity of the propulsive waveform, subsequently impacting UUS forward motion [25, 53,54,55], where shape refers to the frequency, amplitude and temporal coupling of the undulatory movements [16, 54]. Research conducted by Ikeda et al. [41] reported that the movement of the shoulder, lower trunk and lower leg along the body wave was associated with maximum horizontal UUS velocity. As body wave velocity and vertical toe velocity both have a level of independence and are individually related to UUS velocity, optimal UUS performance will only be achieved if the athlete is able to produce high body wave and vertical toe velocity [15].

Hydrodynamics and UUS Parameters

Vortices are integral to the production of propulsive force and minimisation of drag during the UUS [33, 50, 53]. Vortices are described as rotating masses of water that are created by the heaving and pitching motions (cyclic motion) of the toes during UUS [25, 50]. Vortices represent the transfer of momentum from the water to the body, and vice versa, resulting in body acceleration [10, 56]. Higher vertical toe and body wave velocity and vertical toe velocity have been linked to vortex creation and thrust production, with these vortices having the potential to explain some of the differences in athlete’s UUS velocity [10, 15, 56]. The flow pattern behind the swimmer in the wake shows a jet stream between two counter-rotational vortices following the down kick [7, 44]. Similar results were found by Hochstein et al. [44], Miwa et al. [46] and Shimojo et al. [38], who confirmed the existence of a pair of vortices and jet flow in the wake of the swimmer, which are assumed to be a part of a vortex ring or structure that drives propulsion. The majority of propulsive force during the UUS was shown to be generated at the end of the down kick, where swimmers externally rotated their lower limbs and moved their feet closer together [38, 57]. This action subsequently generates a strong cluster of vortices and jet flow, increasing the water momentum in the wake, resulting in thrust [38, 44].

Kinanthropometry and UUS Parameters

Kick amplitude reflects the actions of the hip, knee and ankle, where a larger range of motion at these joints creates bigger counter-rotation vortices that maximise leg propulsion to drive forward motion [14]. The results of the current systematic review reflect a positive relationship between maximum knee (r = 0.88), ankle (r = 0.67) and chest (r = 0.52) flexion and extension angles and UUS velocity, with results for the hip not reported [10]. Optimisation simulations of UUS predict that swimmers who exhibited greater similarity between maximum joint range of motion, specifically extension and flexion angles would have higher average horizontal velocity [58]. This was evidenced in the study conducted by Ikeda et al. [41], where a greater shoulder angle was significantly correlated with UUS velocity (r = 0.80), as this improves streamline position and decreases effects of drag. Similarly, Atkison et al. [10] observed the importance of high levels of upper thoracic flexibility for UUS velocity, as this flexibility is required to dampen body undulations of the lower segments and reduce resistive drag by maintaining a small angle of attack of the arms. Wang [39] suggests that the angular velocity of the trunk may also contribute to the higher kick frequency seen in skilled swimmers, describing how the undulatory wave that is initially observed at the lower trunk may enhance peak angular velocities of the lower limb joints. Ikeda et al. [41] also observed that faster swimmers moved their lower trunk with greater angular displacement, which further increased the angular displacement of the shoulder, knee and lower leg. The results suggest that greater angular displacement of the lower legs and feet may produce thrust and in turn propulsion [41]. However, if such movements become too large, this may lead to greater water resistance due to greater frontal projected area [41].

UUS velocity might be achieved in multiple ways, at one extreme, through large undulatory movements (higher amplitude) to maximise propulsive impulse production with high active drag (high energy requirement/cost) or small movements (smaller amplitude) that produce a reduced amount of propulsive impulse but minimise drag [34]. Hochstein and Blickhan [17] reported that slower swimmers utilised higher kick amplitudes and initiated the undulatory wave at the hands. In the slower swimmers, the larger mass of water being propelled by a higher kick amplitude was negated by a lower kick frequency and increasing drag, which resulted in reduced overall velocity. Consistent with this view, Elipot et al. [14] have found that higher-level swimmers utilise strong joint synergy between the hip, knee and ankle joint, whereby they adopt a regulation loop in which the hip-ankle and the knee action are independent and control the kick amplitude. Skilled athletes appear able to better maximise propulsive impulse during the UUS by employing optimal amplitudes of the end effector that in turn minimise flow separation and maximise energy reuse from the vortices in the wake around the body [59,60,61]. How variations in a swimmer’s kinanthropometric and technical characteristics may influence the combination of joint movements that optimise the relationship between kick amplitude and kick frequency and their resulting UUS velocity remains unclear.

While the coordination of multiple body joints is important to optimise the undulatory movements that maximise UUS velocity, the swimmers’ flexibility across these joints has the potential to influence the swimmers’ coordination patterns, efficiency of thrust production and ability to minimise resistance forces [24]. However, the optimal ranges of motion across multiple joints for maximising UUS velocity still remain unclear [24, 62,63,64,65].

Few studies have investigated the relationship between ankle flexibility and UUS performance, with the results being somewhat equivocal [24, 48]. When ankle plantarflexion range was restricted by tape, it had a negative impact on kick frequency, movement efficiency and UUS velocity [24]. Similarly, Shimojo et al. [48] found that when restricting ankle plantarflexion range, there was a significant decrease in velocity due to the inhibition of the rotational function; however, Froude efficiency (a measure of swimming efficiency) remained unchanged. The hydrodynamic force acting on the foot during active range of motion was thought to be higher than expected as the maximal plantarflexion angle did not decrease following tape application [48]. Shimojo et al. [48] concluded that foot rotational ability, rather than ankle flexibility is associated with increased UUS velocity, appearing consistent with the results of Willems et al. [24].

There was no significant correlation between dryland ankle flexibility measures and UUS velocity [24]. Alternatively, Willems et al. [24] observed positive correlations between dorsiflexion strength (r = 0.53) and ankle internal rotation strength (r = 0.47) to UUS velocity. While the ankle internal rotation strength correlation was expected, the importance of dorsiflexion was unexpected as ankle plantar flexion produces propulsive force to drive thrust and momentum [10]. However, dorsiflexion strength may also play a role during the down kick to maintain a rigid foot position while applying force to the water [10]. Yamakawa et al. [49] suggested that the internal/external rotations of the hip joint may contribute to the control of the direction of the dorsal side of the ankle and foot, therefore identifying this as an area of interest in understanding the UUS.

Optimising Kick Frequency and Amplitude

A number of studies have looked to examine how acute changes in kick frequency may influence UUS velocity [16, 34]. To increase their kick frequency beyond their preferred frequency, swimmers must generate larger torque power and/or reduce resistance forces [16]. As increasing joint torques would require more internal work [66], swimmers may not be able to maintain their preferred kick amplitude when increasing their kick frequency for long periods, resulting in either maintenance or reduction in UUS velocity [16]. Such results appear consistent with the literature that higher swim velocities are associated with increased kick frequency and a maintenance of kick amplitude [16, 19, 36, 37, 67]. For undulatory swimming with preferred kick frequency, the amplitude in the end effector is about one fifth of the body length [21]. Hochstein et al. [21] reported that this amplitude range of 0.2 – 0.3 body lengths can be considered a fixed (physical) constraint, regardless of body size, shape or movement.

In their regression analyses, Connaboy et al. [34] observed that the reduction in explained variance with the removal of participant as the fixed factor may be indicative of the number of possible solutions to the task (maximise UUS velocity) in relation to the individual’s own organism constraints [68]. These emphasise that an athlete’s optimal movement combination when performing UUS may be somewhat different to others, as each swimmer needs to exploit their own organism constraints to maximise propulsive impulse while simultaneously reducing drag in response to the task and environmental constraints [68]. For example, Connaboy et al. [34] reported that in a sample of 17 national swimmers (8 M, 9F), two swimmers had identical mean maximal swimming velocity (1.18 m.s−1). Swimmer A had the lowest kick amplitude of the entire group (0.52 m), but the second highest kick frequency (2.22 Hz). In contrast, Swimmer B had the second highest kick amplitude (0.69 m) but third lowest kick frequency (1.98 Hz). These results provide some preliminary data that swimmers who possess different organism-level constraints, e.g. force production or range of motion capabilities at specific joints, or limb and body length may utilise different movement solutions when performing the UUS [34]. Future research should therefore examine how a combination of factors such as a swimmer's kinanthropometry and technique as well as the swimming stroke and race distance may also alter the movement solutions generated to optimise UUS velocity [21, 34].

Influence of Body Position on UUS Performance Variables

The UUS is a skill utilised in all swimming strokes, with the exclusion of breaststroke. The prone body position is utilised in the front crawl and butterfly, with the dorsal body position seen in backstroke. Prone UUS velocity was highly correlated with kick frequency (r = 0.90) [6] and down kick acceleration decreased significantly during a dorsal body position. During prone and dorsal UUS, a qualitative analysis of the kicking path showed near identical trajectories (amplitude and timing), as well as up and down kick duration [33]. The two studies comparing UUS performance in the dorsal prone position reported minor differences in the UUS velocity, kick amplitude and frequency in six junior national swimmers (17.02 ± 0.36 years) [6] and 11 male national swimmers (19.9 ± 2.15 years) [33]. Similar results were also obtained for kick length and mean body speed between prone and dorsal UUS [33]. Towards the end of knee flexion in both body conditions (the up kick for prone UUS and down kick for dorsal UUS), there were significant differences in the shoulder (P: 163.90°, D: 159.36°, p < 0.05) and knee angle (P: 118.27°, D: 114.27°, p < 0.05), where there was a larger shoulder and knee angle during the prone position [33]. Dorsal kicking seemed to require greater levels of plantarflexion [6], and Arellano et al. [33] described UUS in this body position to have more pronounced body oscillations than in the prone position.

When the swimmer adopts a non-prone UUS position (i.e. dorsal or lateral), body undulations are more evident, which may imply a difference in kicking technique, propulsive force production, joint range of motion and/or resistance forces encountered. This could be related to the prone body position providing a more stable position for the swimmer’s centre of mass compared to the dorsal body position, which may contribute to a more efficient movement pattern [6]. Theoretically, the prone position may allow greater UUS velocity to be generated as a result of its higher kick amplitudes and frequencies and/or reduced resistance forces than other swimming positions. While such results have been shown when comparing prone and dorsal positions to a lateral position, only relatively minor differences have been discovered between prone and dorsal positions [6, 33].

The Impact of Horizontal Distance and Depth on UUS Variables

When entering the water after a dive start, the swimmer’s velocity (~ 3.61 m.s−1) is greater than at any other time of the race due to the resistance of the water being greater than that of air [35, 69]. de Jesus et al. [35] reported that UUS velocities decreased from the first four kick cycles to the final four before surfacing, with this associated with a reduced kick amplitude and frequency. Determining when and where the underwater swimmer should begin kicking compared to maintaining a streamlined glide position is an important practical question [35].

Lyttle and Blanksby [42] suggest that swimmers should perform the glide at approximately 0.4 m underwater at all velocities above 1.9 ms−1 to gain maximum drag reduction benefits, where a 15–18% reduction in total drag was found when compared to swimming at the surface. It has been proposed that propulsive movements should be initiated when the underwater velocity drops to 1.9–2.2 m.s−1 because this is the maximum range of speeds that produced a significant reduction in net force in the kicking conditions compared to streamlined positions [43]. Using this recommendation, Elipot et al. [14] and Houel and colleagues [37] suggest that high-performance swimmers should initiate undulatory movements, using the legs to generate propulsion when the centre of mass is ~ 6 m from the starting wall. At this distance, mean velocities were 1.99 ± 0.13 m.s−1, aligning with findings of Lyttle et al., [43] and Elipot et al. [14]. Houel and Elipot [37] conclude from their results of a multiple stepwise regression and comparison to results of other studies when a swimmer initiates the UUS after completing the glide phase, he/she can maximise their UUS velocity by progressively increasing kick frequency and maintaining their kick amplitude [35].

Methodological Considerations

As with any systematic review or meta-analysis, there are a number of strengths and limitations that may affect the generalisability of the results. These strengths and limitations are summarised with respect to the review methodology as well as the reviewed literature.

Review Methodology

The primary strengths of the review were the pre-registration of the review methodology, the use of the PRISMA methodology and use of a risk of bias tool recently developed for cross-sectional biomechanical studies [1]. The primary limitation of the review process was the restriction of articles to those written in English and that were peer-reviewed journal articles or longer conference proceedings articles. Only articles about human swimmers were included which may limit our understanding of the UUS as other studies involving animals or computational simulation approaches may also have provided some insight into improving UUS velocity in human swimmers.

Measurement of UUS

In terms of the literature, the major strength of this review compared to that of the previous review by Connaboy et al. [34] is the identification of an additional 17 studies. The 23 studies included in the current systematic review exhibited some degree of inconsistency with the measurement of UUS kinematic measures, specifically the distances of the underwater phase (10 m, 15 m, 20 m or 25 m) and starting positions of each trial. The majority of studies [10, 16, 17, 24, 33, 35, 36, 40, 47] used a push start to 10 m or 15 m; however, some studies analysed the UUS following a block start [6, 14, 36]. As a result, determining the correlation of biomechanical parameters to underwater velocity may be influenced by the studies' measurement distance and their approach to determining UUS performance.

There also appear to be some differences in the nature of the swim task performed across the included studies. Within this review, the majority of the studies tested the swimmers in a competition swimming pool and had them perform UUS to a set distance [6, 7, 10, 14,15,16,17, 24, 33,34,35,36,37, 40, 41, 47, 48]. Other studies looked at the hydrodynamics affecting UUS as well as the kinematics and kinetics using a towing mechanism [42, 43] or had the swimmers performing UUS in a swimming flume at different speeds [17, 34, 38, 46]. There was considerable between-study variation in the outcome measures assessed and the data collection and analysis procedures. It is also possible that the variety of UUS methodologies used may have had significant implications in the comparison of results between studies.

Study Population

The between study differences in kinematic, kinetic and kinanthropometric characteristics may have also been influenced by differences in sex, age and performance level. The majority of studies reviewed generally consisted of a small sample size and a potentially greater bias towards male compared to female swimmers. Only two of the observational studies consisted entirely of female swimmers, and the seven studies that had a mix of females and males typically had a greater number of male participants compared to females with the exception of two studies [17, 34]. In addition, the majority of studies did not provide very clear descriptions of the level of swimmer, which may be as a result of an inconsistency in the description of different levels of swimmers and lack of standardisation across research. As such, it is difficult to determine how sex, age and performance level may influence UUS velocity and the kinematics and kinetics of this movement.

Study Design

The smaller sample sizes seen in the majority of the included studies suggest relatively broad confidence intervals for the outcomes as well as the correlations between outcomes reported in many of the studies. Therefore, there may be some uncertainty regarding what constitutes representative values for different levels of swimmers as well as the strength of correlations between these variables. The small sample sizes may have also contributed to relative lack of statistical power when comparing between groups or conditions, which may have meant some true significant differences were not observed. Further, there is no evidence of any randomised controlled trials conducted in the literature, limiting any understanding of true cause-and-effect relationships between variables.

Across the 25 studies included in this systematic review, there appears to be an absence of long-term monitoring or interventional research that has sought to determine the chronic effect of any biomechanically informed skill acquisition or strength and conditioning intervention. This lack of longitudinal studies is a major limiting factor within this literature.

Conclusion

The current systematic review has identified that kick frequency [6, 7, 37], kick amplitude [36, 37], vertical toe velocity [10, 15] and knee angular velocity [15, 34] may be the greatest predictors of UUS velocity. Practical implications of how to optimise these parameters for swimmers with different kinanthropometric characteristics are not yet known as an increase in one characteristic, e.g. kick frequency may actually reduce kick amplitude and potentially UUS velocity.

Due to the greater magnitude of water compared to air resistance, the swimmer will lose horizontal velocity after striking the water. It was observed that UUS velocity, kick frequency and amplitude tend to decrease over the 15 m underwater phase with the initiation of the kick recommended to start at the 6 m mark [36, 37]. Additionally, during the underwater phase of any transition from the wall, maintaining kick amplitude and increasing kick frequency, specifically as the swimmer approaches the surface, will achieve optimal swimming velocity and overall underwater performance [35,36,37]. Despite the advancement of observational studies and development of an understanding of the determinants that impact UUS performance, there appears to be almost a complete lack of intervention research in the biomechanics of UUS. As a result, there is very limited knowledge on how changes in these biomechanical factors interact during performance and why there is variation across different swimmers. Future research should further examine how to optimise the kinematic and kinetic characteristics with respect to the imposed organism constraints [34], as well as the effect of different biomechanically informed skill acquisition and strength and conditioning interventions to improve aspects of UUS performance. Standardisation of testing methods and outcomes related to UUS would also be beneficial.

Availability of Data and Materials

The search term strategy used for each database is provided in the methods section of this systematic review. Data sharing is not applicable to this publication as no datasets were produced or analysed during the study.

Code Availability (Software Application or Custom Code)

Not applicable.

Abbreviations

BSFE:

Backstroke start with feet emerged

BSFI:

Backstroke start with feet immerged

D:

Dorsal

DK:

Down kick

F:

Female

FINA:

Fédération internationale de natation

M:

Male

MeSH:

Medical Subject Headings

P:

Prone

PICO:

Population, intervention, comparison/control and outcome

PRISMA:

Preferred Reporting Items for Systematic Reviews and Meta-analyses

ROM:

Range of motion

SD:

Standard deviation

UK:

Up kick

UUS:

Undulatory underwater kick

References

  1. Hindle BR, Lorimer A, Winwood P, Keogh JWL. The biomechanics and applications of strongman exercises: a systematic review. Sports Med Open. 2019;5(1):49. https://doi.org/10.1186/s40798-019-0222-z.

    Article  PubMed  PubMed Central  Google Scholar 

  2. Domínguez R, Jesús-Sánchez-Oliver A, Cuenca E, Jodra P, FernandesdaSilva S, Mata-Ordóñez F. Nutritional needs in the professional practice of swimming: a review. J Exerc Nutr Biochem. 2017;21(4):1–10. https://doi.org/10.20463/jenb.2017.0030.

    Article  Google Scholar 

  3. Shaw G, Boyd KT, Burke LM, Koivisto A. Nutrition for swimming. Int J Sport Nutr Exerc Metab. 2014;24(4):360–72. https://doi.org/10.1123/ijsnem.2014-0015.

    Article  PubMed  Google Scholar 

  4. Thng S, Pearson S, Keogh JWL. Relationships between dry-land resistance training and swim start performance and effects of such training on the swim start: a systematic review. Sports Med. 2019;49(12):1957–73. https://doi.org/10.1007/s40279-019-01174-x.

    Article  PubMed  Google Scholar 

  5. Jones J. Dry land strength and power training to enhance swimming in-water turn performance. School of Medical and Health Sciences. Research Online: Edith Cowan University; 2017. p. 63.

  6. Alves F, Lopes P, Veloso A, Martins-Silva A. Influence of body position on dolphin kick kinematics. In: XXIV international symposium on biomechanics in sports. Salzburg, Austria, University of Salzburg: International Society of Biomechanics in Sports; 2006. p. 67–70.

  7. Hochstein S, Blickhan R. Vortex re-capturing and kinematics in human underwater undulatory swimming. Hum Mov Sci. 2011;30(5):998–1007.

    Article  Google Scholar 

  8. Takeda T, Sakai S, Takagi H. Underwater flutter kicking causes deceleration in start and turn segments of front crawl. Sports Biomech. 2020. https://doi.org/10.1080/14763141.2020.1747528.

    Article  PubMed  Google Scholar 

  9. Natation FID: Fina Swimming Rules. 2017. https://resources.fina.org/fina/document/2021/01/12/b3885f9b-630a-4390-861d-4e7f6031f4a4/2017_2021_swimming_16032018.pdf.

  10. Atkison RR, Dickey JP, Dragunas A, Nolte V. Importance of sagittal kick symmetry for underwater dolphin kick performance. Hum Mov Sci. 2014;33:298–311.

    Article  Google Scholar 

  11. Cortesi M, Gatta G. Effect of the swimmer’s head position on passive drag. J Hum Kinet. 2015;49:37–45. https://doi.org/10.1515/hukin-2015-0106.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Vennell R, Pease D, Wilson B. Wave drag on human swimmers. J Biomech. 2006;39(4):664–71. https://doi.org/10.1016/j.jbiomech.2005.01.023.

    Article  PubMed  Google Scholar 

  13. Lyttle A, Blanksby B, Elliott B, Lloyd D. A comparison of underwater gliding and kicking techniques. In: Sanders RHG, editors. XVII international symposium on biomechanics in sports. Edith Cowan University, Perth, Western Australia: Perth School of Biomedical and Sports Science, Edith Cowan University; 1999. p. 81–4.

  14. Elipot M, Dietrich G, Hellard P, Houel N. High-level swimmers’ kinetic efficiency during the underwater phase of a grab start. J Appl Biomech. 2010;26(4):501–7.

    Article  Google Scholar 

  15. Higgs AJ, Pease DL, Sanders RH. Relationships between kinematics and undulatory underwater swimming performance. J Sports Sci. 2017;35(10):995–1003. https://doi.org/10.1080/02640414.2016.1208836.

    Article  PubMed  Google Scholar 

  16. Shimojo H, Sengoku Y, Miyoshi T, Tsubakimoto S, Takagi H. Effect of imposing changes in kick frequency on kinematics during undulatory underwater swimming at maximal effort in male swimmers. Hum Mov Sci. 2014;38:94–105. https://doi.org/10.1016/j.humov.2014.09.001.

    Article  PubMed  Google Scholar 

  17. Hochstein S, Blickhan R. Body movement distribution with respect to swimmer’s glide position in human underwater undulatory swimming. Hum Mov Sci. 2014;38:305–18. https://doi.org/10.1016/j.humov.2014.08.017.

    Article  PubMed  Google Scholar 

  18. Sanders RH, Cappaert JM, Devlin RK. Wave characteristics of butterfly swimming. J Biomech. 1995;28(1):9–16. https://doi.org/10.1016/0021-9290(95)80002-6.

    Article  CAS  PubMed  Google Scholar 

  19. Gavilán A, Arellano R, Sanders R. Underwater undulatory swimming: study of frequency, amplitude and phase characteristics of the “body wave.” Port J Sport Sci Suppl Biomech Med Swim X. 2006;6:35–7.

    Google Scholar 

  20. McHenry MJ, Pell CA, Long JH Jr. Mechanical control of swimming speed: stiffness and axial wave form in undulating fish models. J Exp Biol. 1995;198(Pt 11):2293–305.

    Article  CAS  Google Scholar 

  21. Hochstein S, Meyer A, Blickhan R. Effect of kick frequency variation on swimming speed and kinematics in human underwater undulatory swimming. In: Exercise JSoSiSaW, editor. Biomechanics and medicine in swimming, XIII. Tokyo: Impress R&D; 2018. p. 71–8.

  22. Pyne DB, Sharp RL. Physical and energy requirements of competitive swimming events. Int J Sport Nutr Exerc Metab. 2014;24(4):351–9.

    Article  Google Scholar 

  23. von Loebbecke A, Mittal R, Fish F, Mark R. A comparison of the kinematics of the dolphin kick in humans and cetaceans. Hum Mov Sci. 2009;28(1):99–112. https://doi.org/10.1016/j.humov.2008.07.005.

    Article  Google Scholar 

  24. Willems TM, Cornelis JA, De Deurwaerder LE, Roelandt F, De Mits S. The effect of ankle muscle strength and flexibility on dolphin kick performance in competitive swimmers. Hum Mov Sci. 2014;36:167–76. https://doi.org/10.1016/j.humov.2014.05.004.

    Article  PubMed  Google Scholar 

  25. Connaboy C, Coleman S, Sanders RH. Hydrodynamics of undulatory underwater swimming: a review. Sports Biomech. 2009;8(4):360–80. https://doi.org/10.1080/14763140903464321.

    Article  PubMed  Google Scholar 

  26. Bingul BM, Tore O, Bulgan C, Aydin M. The kinematic analysis of the grab, rear track and front track start in swimming. Sport Mont. 2015;43–45:57–62.

    Google Scholar 

  27. Slawson S, Conway P, Cossor J, Chakravorti N, West A. The categorisation of swimming start performance with reference to force generation on the main block and footrest components of the Omega OSB11 start blocks. J Sports Sci. 2013;31(5):468–78.

    Article  Google Scholar 

  28. Tor E, Pease DL, Ball KA. Key parameters of the swimming start and their relationship to start performance. J Sports Sci. 2015;33(13):1313–21. https://doi.org/10.1080/02640414.2014.990486.

    Article  PubMed  Google Scholar 

  29. Page MJ, McKenzie JE, Bossuyt PM, Boutron I, Hoffmann TC, Mulrow CD, et al. The PRISMA 2020 statement: an updated guideline for reporting systematic reviews. J Clin Epidemiol. 2021;134:178–89. https://doi.org/10.1016/j.jclinepi.2021.03.001.

    Article  PubMed  Google Scholar 

  30. Natera AO, Cardinale M, Keogh JWL. The effect of high volume power training on repeated high-intensity performance and the assessment of repeat power ability: a systematic review. Sports Med. 2020;50(7):1317–39. https://doi.org/10.1007/s40279-020-01273-0.

    Article  PubMed  Google Scholar 

  31. Hindle BR, Lorimer A, Winwood P, Keogh JWL. A systematic review of the biomechanical research methods used in strongman studies. Sports Biomech. 2020;19(1):90–119. https://doi.org/10.1080/14763141.2019.1598480.

    Article  PubMed  Google Scholar 

  32. Davids EL, Roman NV. A systematic review of the relationship between parenting styles and children’s physical activity. Afr J Phys Health Educ Recreat Dance. 2014;20(Supplement 2):228–46.

    Google Scholar 

  33. Arellano R, Gavilan A, Garcia F. A comparison of the underwater undulatory swimming technique in two different body positions. In: VIII international symposium on biomechanics and medicine in swimming. University of Jyvaskyla, Finland: Biomechanics and Medicine in Swimming VIII; 1998. p. 25–8.

  34. Connaboy C, Naemi R, Brown S, Psycharakis S, McCabe C, Coleman S, et al. The key kinematic determinants of undulatory underwater swimming at maximal velocity. J Sports Sci. 2016;34(11):1036–43.

    Article  Google Scholar 

  35. de Jesus K, de Jesus K, Machado L, Fernades R, Vilas-Boas JP. Linear kinematic of the underwater undulatory swimming phase performed after two backstroke starting techniques. In: International symposium on biomechanics in sports: conference proceedings archive; 2012. p. 371–4.

  36. Houel N, Elipot M, Andrée F, Hellard H. Kinematics analysis of undulatory underwater swimming during a grab start of national level swimmers. In: XIth international symposium for biomechanics & medicine in swimming. 2010(11), p. 97–9.

  37. Houel N, Elipot M, Andre F, Hellard P. Influence of angles of attack, frequency and kick amplitude on swimmer’s horizontal velocity during underwater phase of a grab start. J Appl Biomech. 2013;29(1):49–54. https://doi.org/10.1123/jab.29.1.49.

    Article  PubMed  Google Scholar 

  38. Shimojo H, Gonjo T, Sakakibara J, Sengoku Y, Sanders R, Takagi H. A quasi three-dimensional visualization of unsteady wake flow in human undulatory swimming. J Biomech. 2019;93:60–9. https://doi.org/10.1016/j.jbiomech.2019.06.013.

    Article  PubMed  Google Scholar 

  39. Wang N, Liu Y. Kinetic chain application to the dolphin kick in butterfly swimming. In: International society of biomechanics in sports, proceedings of XXIV international symposium on biomechanics in sports 2006, Salzburg, Austria, University of Salzburg, c2006; 2006. p. 469–71.

  40. Yamakawa KK, Shimojo H, Takagi H, Tsubakimoto S, Sengoku Y. Effect of increased kick frequency on propelling efficiency and muscular co-activation during underwater dolphin kick. Hum Mov Sci. 2017;54:276–86. https://doi.org/10.1016/j.humov.2017.06.002.

    Article  PubMed  Google Scholar 

  41. Ikeda Y, Ichikawa H, Shimojo H, Nara R, Baba Y, Shimoyama Y. Relationship between dolphin kick movement in humans and velocity during undulatory underwater swimming. J Sports Sci. 2021;25:1–7. https://doi.org/10.1080/02640414.2021.1881313.

    Article  Google Scholar 

  42. Lyttle A, Blanksby B. A look at gliding and underwater kicking in the swim turn. In: Sanders RHY, editor. XVIII international symposium on biomechanics in sports. The Chinese University of Hong Kong, Hong Kong: Applied program: application of biomechanical study in swimming; 2000. p. 56–63.

  43. Lyttle AD, Blanksby BA, Elliott BC, Lloyd DG. Net forces during tethered simulation of underwater streamlined gliding and kicking techniques of the freestyle turn. J Sports Sci. 2000;18(10):801–7. https://doi.org/10.1080/026404100419856.

    Article  CAS  PubMed  Google Scholar 

  44. Hochstein S, Pacholak S, Brücker C, Siebert T, Blickhan R. Drag reduction by underwater undulatory swimming? An experimental and numerical approach. In: XIIth international symposium on biomechanics and medicine in swimming. Canberra: Australian Institute of Sport; 2014. p. 141–7.

  45. Jensen RK, McIlwain J. Modeling lower extremity forces in the dolphin kick. In: Terauds J, Bedingfield EW, editors. Swimming III, Baltimore. University Park: University Park Press; 1979. p. 137–47.

    Google Scholar 

  46. Miwa T, Matsuuchi K, Shintani H, Kamata E, Nomura T. Unsteady flow measurement of dolphin kicking wake in sagittal plane using 2C-PIV. In: Xth congress of biomechanics and medicine in swimming. Portuguese Journal of Sport Sciences; 2006. p. 64–6.

  47. Shimojo H, Ichikawa H, Tsubakimoto S, Takagi H. The effect of a target sound made by a model swimmer's dolphin kick movement on another swimmer's dolphin kick performance. In: XIth international symposium for biomechanics & medicine in swimming; 2010. p. 341–3.

  48. Shimojo H, Nara R, Baba Y, Ichikawa H, Ikeda Y, Shimoyama Y. Does ankle joint flexibility affect underwater kicking efficiency and three-dimensional kinematics? J Sports Sci. 2019;37(20):2339–46.

    Article  Google Scholar 

  49. Yamakawa KK, Takagi H, Sengoku Y. Three-dimensional analysis of hip and knee joint movements during dolphin kicking and butterfly swimming. In: XIII biomechanics and medicine in swimming. Tokyo: Impress R&D; 2018. p. 187–92.

  50. Arellano R, Pardillo S, Gavilán A. Underwater undulatory swimming: Kinematic characteristics, vortex generation and application during the start, turn and swimming strokes. In: Proceedings of the XXth international symposium on biomechanics in sports. 2002.

  51. Matsuuchi K, Hashizume T, Nakazawa Y, Nomura T, Shintani H, Miwa T. Flow visualization of unsteady flow field around a monofin using PIV. X Biomech Med Swim. 2006;6(2 Suppl):60–2.

    Google Scholar 

  52. Bressel E. Innovative analyses of human movement: analytical tools for human movement research. Med Sci Sports Exerc. 2004;36(10):1834.

    Article  Google Scholar 

  53. Connaboy C, Coleman S, Moir G, Sanders R. Measures of reliability in the kinematics of maximal undulatory underwater swimming. Med Sci Sports Exerc. 2010;42(4):762–70. https://doi.org/10.1249/MSS.0b013e3181badc68.

    Article  PubMed  Google Scholar 

  54. Castell S, Lighthill J. Mathematical biofluid dynamics. Math Gaz. 1976;60:67. https://doi.org/10.2307/3615654.

    Article  Google Scholar 

  55. Videler J, Kamermans P. Differences between upstroke and downstroke in swimming dolphins. J Exp Biol. 1985;119:265–74.

    Article  CAS  Google Scholar 

  56. Ungerechts B, Persyn U, Colman V. Analysis of swimming techniques using vortex traces. In: 18 international symposium on biomechanics in sports. Hong Kong: ISBS—conference proceedings archive; 2000.

  57. Cohen RC, Cleary PW, Mason BR. Simulations of dolphin kick swimming using smoothed particle hydrodynamics. Hum Mov Sci. 2012;31(3):604–19. https://doi.org/10.1016/j.humov.2011.06.008.

    Article  PubMed  Google Scholar 

  58. Nakashima M. Simulation analysis of the effect of trunk undulation on swimming performance in underwater dolphin kick of human. J Biomech Sci Eng. 2009;4:94–104. https://doi.org/10.1299/jbse.4.94.

    Article  Google Scholar 

  59. Tokumaru PT, Dimotakis PE. Rotary oscillation control of a cylinder wake. J Fluid Mech. 1991;224:77–90. https://doi.org/10.1017/S0022112091001659.

    Article  Google Scholar 

  60. Triantafyllou M, Techet A, Zhu Q, Beal D, Hover F, Yue D. Vorticity control in fish-like propulsion and maneuvering. Integr Comp Biol. 2002;42:1026–31. https://doi.org/10.1093/icb/42.5.1026.

    Article  CAS  PubMed  Google Scholar 

  61. Streitlien K, Barrett D, Triantafyllou M. Oscillating foils of high propulsive efficiency. J Fluid Mech. 1998;360:41–72. https://doi.org/10.1017/S0022112097008392.

    Article  Google Scholar 

  62. Jagomägi G, Jürimäe T. The influence of anthropometrical and flexibility parameters on the results of breaststroke swimming. Anthropol Anz. 2005;63(2):213–9.

    Article  Google Scholar 

  63. McCullough AS, Kraemer WJ, Volek JS, Solomon-Hill GF Jr, Hatfield DL, Vingren JL, et al. Factors affecting flutter kicking speed in women who are competitive and recreational swimmers. J Strength Cond Res. 2009;23(7):2130–6. https://doi.org/10.1519/JSC.0b013e31819ab977.

    Article  PubMed  Google Scholar 

  64. Sugimoto S, Nakashima M, Ichikawa H, Miwa T, Takeda T. The effects of plantar flexion angle increment on the performance during underwater dolphin kick using simulation analysis. Jpn J Phys Educ Health Sport Sci. 2008;53:51–60. https://doi.org/10.5432/jjpehss.0522.

    Article  Google Scholar 

  65. Kippenhan B. Lower-extremity joint angles used during the breaststroke whip kick. In: Proceedings of swim sessions: XIX international symposium on biomechanics in sports. San Francisco;2001. p. 48–52.

  66. Zamparo P, Pendergast DR, Termin B, Minetti AE. How fins affect the economy and efficiency of human swimming. J Exp Biol. 2002;205(Pt 17):2665–76.

    Article  CAS  Google Scholar 

  67. Von Loebbecke A, Mittal R, Fish F, Mark R. Propulsive efficiency of the underwater dolphin kick in humans. J Biomech Eng. 2009;131(5):5. https://doi.org/10.1115/1.3116150.

    Article  Google Scholar 

  68. Newell KM. Constraints on the development of coordination. In: Wade MG, editor. Motor development in children: aspects of coordination and control. The Netherlands: Martinus Nijhoff, Dordrecht; 1986. p. 341–60.

  69. Elipot M, Hellard P, Taïar R, Boissière E, Rey JL, Lecat S, et al. Analysis of swimmers’ velocity during the underwater gliding motion following grab start. J Biomech. 2009;42(9):1367–70. https://doi.org/10.1016/j.jbiomech.2009.03.032.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors wish to acknowledge Mr David Honeyman for his help during the database search process and Mr Oran Mcinerney for his contribution during the screening of the literature and quality/risk of bias assessment process.

Funding

This research is supported by an Australian Government Research Training Program Scholarship and The Queensland Academy of Sport, Sport Performance Innovation and Knowledge Excellence to the primary author. However, no sources of funding were used in the preparation of this article.

Author information

Authors and Affiliations

Authors

Contributions

RW, AL, SP and JK all contributed to the planning and design of the systematic review. RW was responsible for the screening and quality/risk of bias assessment of literature. RW was responsible for data extraction. RW, AL, SP and JK analysed the data and drafted the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Rani West.

Ethics declarations

Ethical Approval and Consent to Participate

Not applicable.

Consent for Publication

Not applicable.

Competing interests

Rani West, Anna Lorimer, Simon Pearson and Justin Keogh declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

West, R., Lorimer, A., Pearson, S. et al. The Relationship Between Undulatory Underwater Kick Performance Determinants and Underwater Velocity in Competitive Swimmers: A Systematic Review. Sports Med - Open 8, 95 (2022). https://doi.org/10.1186/s40798-022-00485-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40798-022-00485-0

Keywords