Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Assessment of Model Based (Input) Impedance, Pulse Wave Velocity, and Wave Reflection in the Asklepios Cohort

  • Bernhard Hametner ,

    Contributed equally to this work with: Bernhard Hametner, Siegfried Wassertheurer

    Affiliation Health & Environment Department, AIT Austrian Institute of Technology, Vienna, Austria

  • Stephanie Parragh,

    Affiliations Health & Environment Department, AIT Austrian Institute of Technology, Vienna, Austria, Department of Analysis & Scientific Computing, Vienna University of Technology, Vienna, Austria

  • Christopher Mayer,

    Affiliation Health & Environment Department, AIT Austrian Institute of Technology, Vienna, Austria

  • Thomas Weber,

    Affiliation Cardiology Department, Klinikum Wels-Grieskirchen, Wels, Austria

  • Luc Van Bortel,

    Affiliation Department of Pharmacology, Ghent University, Ghent, Belgium

  • Marc De Buyzere,

    Affiliation Department of Pharmacology, Ghent University, Ghent, Belgium

  • Patrick Segers,

    Affiliation Institute of Biomedical Technology, iMinds Medical IT, Ghent University, Ghent, Belgium

  • Ernst Rietzschel,

    Affiliation Department of Internal Medicine, Ghent University, Ghent, Belgium

  • Siegfried Wassertheurer

    Contributed equally to this work with: Bernhard Hametner, Siegfried Wassertheurer

    Siegfried.Wassertheurer@ait.ac.at

    Affiliations Health & Environment Department, AIT Austrian Institute of Technology, Vienna, Austria, Department of Analysis & Scientific Computing, Vienna University of Technology, Vienna, Austria

Abstract

Objectives

Arterial stiffness and wave reflection parameters assessed from both invasive and non-invasive pressure and flow readings are used as surrogates for ventricular and vascular load. They have been reported to predict adverse cardiovascular events, but clinical assessment is laborious and may limit widespread use. This study aims to investigate measures of arterial stiffness and central hemodynamics provided by arterial tonometry alone and in combination with aortic root flows derived by echocardiography against surrogates derived by a mathematical pressure and flow model in a healthy middle-aged cohort.

Methods

Measurements of carotid artery tonometry and echocardiography were performed on 2226 ASKLEPIOS study participants and parameters of systemic hemodynamics, arterial stiffness and wave reflection based on pressure and flow were measured. In a second step, the analysis was repeated but echocardiography derived flows were substituted by flows provided by a novel mathematical model. This was followed by a quantitative method comparison.

Results

All investigated parameters showed a significant association between the methods. Overall agreement was acceptable for all parameters (mean differences: -0.0102 (0.033 SD) mmHg*s/ml for characteristic impedance, 0.36 (4.21 SD) mmHg for forward pressure amplitude, 2.26 (3.51 SD) mmHg for backward pressure amplitude and 0.717 (1.25 SD) m/s for pulse wave velocity).

Conclusion

The results indicate that the use of model-based surrogates in a healthy middle aged cohort is feasible and deserves further attention.

Introduction

The consequences of arterial stiffening on aortic hemodynamics and left ventricular load are actually considered to be major determinants of cardiovascular risk beyond established risk factors [1,2]. Starting from the late 1960’s, methods were established to quantify arterial stiffness and wave reflection at the proximal aorta under different conditions based on the concept of vascular impedance [3,4]. Research was mainly driven by invasive data assessment in the early days, but the technical development of non-invasive sensor systems like Doppler ultrasound and vascular tonometry of superficial arteries facilitated non-invasive data capturing [5]. For example, carotid to femoral pulse wave velocity measurement evolved using these techniques and proofed its clinical usefulness [6,7]. Nevertheless, the acquisition of pulse waves or left ventricular outflow remains time consuming and requires skilled operators as well as dedicated devices. Subsequently, mathematical methods were investigated to substitute measured flow waveforms by approximate or model-predicted alternatives [8,9]. Based on this simplification, new research applications became feasible and parameters based on non-invasively assessed pressure alone turned out to predict cardiovascular events in different cohorts independent of established risk factors [1012]. Such mathematical models potentially allow the calculation of surrogates of characteristic impedance (Zc) [9,13,14]. Zc is a descriptor of combined geometrical and mechanical arterial wall properties, a major determinant of pulse wave velocity, and directly linked to vascular load [1517]. A recently introduced model-based approach, which is grounded on a modified Windkessel system, aims to allow the calculation of input as well as characteristic impedance, pulse wave velocity and wave reflections from pressure waveforms alone [18]. The aim of this work is the comparison of Windkessel derived parameters against directly measured values in an independent large cohort (from the ASKLEPIOS study [19,20]).

Methods

Study population

All data was selected from the ASKLEPIOS study which is a prospective longitudinal study intended to investigate the development of cardiovascular disease in the general population. The actual study population used here includes 1163 women and 1063 men within an age range from 35 to 56 years. The study was conducted on-site in Erpe Mere, Belgium, with approval of the ethics committee of the Ghent University Hospital. All subjects agreed to join the study by written informed consent. Rationale, design, methods and baseline characteristics of the ASKLEPIOS study have been extensively published elsewhere [19]. Therefore, only a very brief methods summary will be given in the following paragraph and basic characteristics in Table A in S1 File. All parameters assessed by the methods described in the ASKLEPIOS study outline will be referred to as “Asklepios” throughout the manuscript.

Data assessment

Aortic flow waveforms were captured by Echo-Doppler measurements and stroke volume subsequently from the cross sectional area of the left ventricular outflow tract by a Vivid7 ultrasound machine (GE Vingmed Ultrasound). Carotid pulse waveform readings were performed by vascular tonometry of the left common carotid artery. Absolute pressure calibration was done by brachial mean and diastolic pressure as described in [21]. All measurements were performed consecutively by a single, well trained operator (E.R.). Input impedance was calculated following Fourier decomposition of an averaged pressure and flow waveform, and defined as the ratio of corresponding harmonics of pressure and flow. Zc was assessed in the frequency domain and calculated as the average of harmonics 3 to 10 with exclusion of outliers [20]. Carotid pulse waveforms (Pc) were separated into their forward (Pf) and backward (Pb) traveling components using wave separation analysis (WSA). The reflection magnitude (RM) is calculated as the ratio of backward and forward wave amplitudes. Systemic vascular resistance (SVR) was set equal to the 0 Hz frequency of input impedance (ratio of mean pressure and flow). Aortic stiffness was assessed by means of carotid-femoral pulse wave velocity (Asklepios cfPWV).

ARCSolver

The ARCSolver (AIT Austrian Institute of Technology, Vienna, Austria) method is intended to mathematically describe left ventricular outflow during systole according to a given pressure wave. The model combines a modified 3-element like Windkessel system and transmission line theory. This approach allows the estimation of static hemodynamic parameters like stroke volume (SV) or systemic vascular resistance (SVR) as well as measures of pulsatile hemodynamics, like Pf or Pb by means of wave separation analysis and PWV as a derivative of Zc. A detailed technical description and validation in the initial cohort [9,18,2224] as well as data on its predictive value is given elsewhere [11,25]. Fig 1 illustrates the basic principle of the method. Basically, the algorithms were designed for the use with aortic waveforms but in this study we utilized carotid waveforms as a surrogate for the first time. To be able to process carotid waveforms of the ASKLEPIOPS cohort we were obliged to slightly adapt the existing modelling approach: To identify the parameters in the Windkessel model, an optimization routine based on the minimization of left ventricular work is used [18]. Here, we modified the initial values for the parameter estimation process, but left the model, all other parameter settings and algorithms unchanged. Compared to known static flow waveform models [26], the Windkessel based ARCSolver aortic flow wave changes depending on arterial compliance and SVR, as shown in Fig 1. No additional anthropometric inputs are needed for the mathematical model for stroke volume and blood flow wave shape, although information on age and sex are used for signal pre- and post-processing as described earlier [18, 22].

thumbnail
Fig 1. Changes in the shape of the modeled flow wave and the resulting stroke volume SV (ml) for a given heart rate HR and blood pressure level depending on arterial compliance Ca (ml/mmHg) and systemic vascular resistance SVR (mmHg*s/ml).

Left: real scale, right: normalized to a height of 1 arbitrary unit (AU). Parameter values for A (Ca = 0.8, SVR = 1.0, SV = 99), B (Ca = 1.0, SVR = 1.4, SV = 71) and C (Ca = 1.2, SVR 1.8, SV = 55).

https://doi.org/10.1371/journal.pone.0141656.g001

Statistics

Unless stated otherwise, results are expressed as mean and standard deviation (SD) in the statistical analysis. Data of the comparisons were analyzed using the method of Bland-Altman [27]. The correlation between variables was calculated using Pearson’s correlation coefficient. To analyze determinants of estimated and measured data, regression analysis was applied. If not stated otherwise, a level of significance of p = 0.05 was used in all tests. Analyses were performed using MedCalc 12.3 (MedCalc software, Mariakerke, Belgium).

Results

Results for the whole study population and for men and women separately are presented in Table 1. Detailed results of subgroup analysis according to age and sex are given in Table B in S1 File. All investigated parameters showed a significant association between the methods. In detail, mean amplitudes of separated forward pressure waves (Pf) are 43.0 (9.48 SD) mmHg and 42.6 (9.28 SD) mmHg for ARCSolver and Asklepios method, respectively, with a mean difference of 0.356 (4.21 SD) mmHg. Correlation is close with Pearson R = 0.90, compare Fig 2A. Mean amplitudes of the backward pressure waves (Pb) are 22.4 (5.44 SD) mmHg and 20.1 (5.22 SD) mmHg with a mean difference of 2.26 (3.51 SD) mmHg. For Pb, correlation is R = 0.78 (Fig 2B). A Bland-Altman analysis for both parameters is shown in Fig 2D and 2E. In a subgroup analysis, similar trends over age for both methods can be seen except for a slight offset in Pb, compare Fig 3A and 3B. The mean ratio or reflection magnitude (RM) of Pb and Pf is 0.529 (0.103 SD) and 0.477 (0.0872 SD) for Windkessel and Doppler based methods respectively with a significant correlation of R = 0.63.

thumbnail
Fig 2. Scatter plots with regression lines and Bland Altman plots comparing the amplitudes of the forward (A, D) and backward (B, E) traveling pressure waves obtained with the ARCSolver method and the Doppler-ultrasound flow, as well as the estimated ARCSolver pulse wave velocity and the measured Asklepios carotid-femoral PWV (C, F).

Bold letters in the regression equations indicate P<0.001.

https://doi.org/10.1371/journal.pone.0141656.g002

thumbnail
Fig 3. Amplitude of the forward Pf and backward Pb traveling pressure wave (A, B) and pulse wave velocity (C) as functions of sex and age.

Q1, 35–40 years, 277/300 (m/w); Q2, 41–45 years, 274/303 (m/w); Q3, 46–50 years 265/281 (m/w); Q4, 51–56 years, 247/279 (m/w).

https://doi.org/10.1371/journal.pone.0141656.g003

thumbnail
Table 1. Comparison of hemodynamic parameters in the total study population and per gender group.

https://doi.org/10.1371/journal.pone.0141656.t001

Mean Zc estimated by the model was 0.0986 (0.0248 SD) mmHg*s/ml compared to 0.109 (0.0377 SD) derived from echocardiography with a mean difference of -0.0102 (0.0330 SD) and a correlation coefficient of R = 0.51. Similar effects for both approaches could be observed for the correlation to age as they were slightly negative (R = -0.07 for ARCSolver, R = -0.12 for Asklepios) as shown in Fig 4. In addition, the negative association was stronger in the male subgroup for both methods (men: R = -0.14 vs. R = -0.16; women: R = -0.02 vs. R = -0.09, ARCSolver vs. Asklepios). Table 2 shows the determinants of characteristic impedance for both approaches as well as for their difference.

thumbnail
Table 2. Analysis of determinants for characteristic impedance for both methods by a linear regression model (stepwise, enter if P<0.05, remove if P>0.1).

https://doi.org/10.1371/journal.pone.0141656.t002

Estimated pulse wave velocity within the cohort was 7.28 (0.93 SD) m/s and carotid to femoral PWV was 6.56 (1.31 SD) m/s with a mean difference of 0.717 (1.25 SD) m/s and R = 0.42. Scatter and Bland-Altman plot analysis (Fig 2C and 2F) shows only a slight although significant trend. Beyond a gender-specific offset, both methods show a similar behavior with regard to age as illustrated in Fig 3C.

Mean stroke volume is 79.4 (14.3 SD) ml for the mathematical model and 77.3 (18.0 SD) ml measured by echocardiography. Derived systemic vascular resistance (SVR) is therefore 1.21 (0.232 SD) mmHg*s/ml for modeled blood flow and 1.28 (0.310 SD) mmHg*s/ml for measured blood flow, respectively.

Discussion

The aim of this work was the comparison between measured surrogates of arterial stiffness and wave reflections and estimated ones derived from a model based approach in the ASKLEPIOS cohort. As indicated by previous work, calculation of wave separation parameters (WSA) seems to be very robust due to their independence of absolute values with regard to Zc [8]. The WSA parameters showed a good agreement over the whole range of age and gender. Mean difference as well as standard deviation of both biomarkers (Pf and Pb) showed only slight trends for the residuals. The modest systematic offset in Pb may be due to the fact that the ARCSolver is designed to work with aortic waveforms which show a less steep upstroke in early systole in contrast to the carotid waveforms used in this study. This offset also caused a systematic overestimation of the reflection magnitude RM by the ARCSolver method compared to the Doppler flow. Thus, a potential correction of Pb will also affect RM positively and further investigations are therefore indicated. Nevertheless, the results already strengthen data on outcome published earlier for model-based methods [1012].

While WSA parameters are supposed to serve as surrogates of pulsatile hemodynamics, cardiac output and systemic vascular resistance are seen to represent the ‘steady’ cardiovascular properties due to their relation to mean arterial pressure. The systemic vascular resistance shows similar trends over age for both methods. In more detail, the increase in SVR with age is more pronounced in women than in men for both methods. This increase over age is slightly more pronounced for the ARCSolver method. SVR is directly linked to stroke volume via mean arterial pressure and heart rate and therefore stroke volume shows a similar behavior in a mirrored way, see Fig 3D. Nevertheless, absolute values of stroke volume remain a challenge for currently available noninvasive measurement techniques.

Characteristic impedance represents the influence of the arterial wall and subsequently the link between pressure and flow. Zc is linked to arterial stiffness, vascular load and anthropometric measures. The direct comparison shows again acceptable agreement. The analysis of determinants for Zc showed similar contribution of carotid pulse pressure, sex and age for both methods but, in contrast to the Doppler based method, no significant influence of body mass index on the model based approach. The influence of height on Zc in both models is only modest and should be interpreted with caution, because the effects of body size are already partly covered by the gender variable. Similar effects were found in a recent study on untreated hypertensive subjects [17], where ARCSolver Zc was not related to BMI or body surface area. However Zc was significantly correlated with relative wall thickness and the left ventricular mass index. Furthermore, a slightly negative correlation with age was observed for Zc for both methods, which was more pronounced in male subjects. Such behavior has also been observed in other cohorts [28,29]. This seems to be a paradox phenomenon at a first glance but may be explained by a potential increase of aortic diameter with age in combination with a more pronounced stiffening of elastic arteries in women than in men in this age range [3032].

Carotid to femoral pulse wave velocity is the actual non-invasive gold standard for the estimation of aortic pulse wave velocity (PWV) [7]. In a recently published meta-analysis PWV could demonstrate its additional predictive value beyond established risk scores for several target groups [6]. To spread its useful application, simplified acquisition approaches may help. The Windkessel based aortic PWV estimation method has already been compared in other cohorts with cfPWV and invasive aortic measurements [22, 23, 33,34]. Observed results were similar with those in this community and rating according to the ARTERY Society recommendations [35] turned out to be equal. In these recommendations, a mean difference < 1m/sec (with a SD of < 1.5 m/sec)–as observed in our study—between the gold standard and the comparator is classified as acceptable. However, further work may be beneficial to verify this agreement over the whole age range. The narrow age window might also explain the moderate correlation between cfPWV and ARCSolver PWV observed in this study, as age is the most important determinant of PWV. In a recent study, ARCSolver PWV from aortic pressure curves was closely related to cfPWV and especially invasive PWV over a wide age range [33]. Furthermore, prospective longitudinal outcome data of a chronic kidney disease stage 2–4 population was published supporting the predictive value associated with single point estimation of PWV by the ARCSolver method [25]. Upcoming assessment of adverse cardiovascular events in this cohort will help to provide further prospective evidence. Fig 3 reveals that cf-PWV in the Asklepios cohort develops similarly over age for men and women, while for ARCSolver PWV higher values for men but a more rapid increase for women can be seen. It remains unclear which progressions reflect true aortic PWV. Vermeersch et al. extensively investigated local and global stiffness behavior in the Asklepios cohort [32]. They found a steeper increase for women compared to men over age for local carotid PWV and higher values in local femoral PWV in men over all age groups. In several studies looking at pulse wave velocity development over age for both sexes (carotid-femoral as well as brachial-ankle PWV), the effect of a lower PWV in women at younger ages but a steeper increase leading to similar or higher PWV in older ages can be seen. These progressions are differently pronounced, depending on the specific study and cohort [3639]. They show sometimes small but sometimes also significant differences.

Limitations

In this study carotid pulse waves were applied to the ARCSolver algorithms instead of central aortic pulse waves for the first time. Therefore, slight adaptions in the signal processing chain were necessary although no modifications in the model as well as the algorithms themselves were performed. Nevertheless, the modifications possibly influence the current findings. Confirmation in other cohorts may be seen as useful. Furthermore, the study population consisted of middle-aged, healthy subjects only, thus results may not be generalizable to other cohorts. Because of the narrow age-range, this may be especially true for PWV.

Conclusion

Overall, we observed acceptable agreement for all studied parameters. The results therefore indicate that the use of model-based surrogates in a healthy middle aged cohort is feasible and deserves further attention. Furthermore, this comparison implicitly also allowed a cross-validation of a potential operator dependent bias with regard to age and sex on cohort level for which no evidence could be found. For both methods, the parameters of arterial function compared here showed their predictive power already in prospective trials and independent cohorts. The presented results show consistency and strengthen previous findings.

Supporting Information

Author Contributions

Conceived and designed the experiments: LVB MDB PS ER. Performed the experiments: ER. Analyzed the data: BH SP PS SW. Wrote the paper: BH SP CM TW PS SW.

References

  1. 1. Mancia G, Fagard R, Narkiewicz K, Redón J, Zanchetti A, Böhm M, et al. 2013 ESH/ESC Guidelines for the management of arterial hypertension: The Task Force for the management of arterial hypertension of the European Society of Hypertension (ESH) and of the European Society of Cardiology (ESC). Eur Heart J. 2013; 34:2159–2219. pmid:23771844
  2. 2. Laurent S, Cockcroft J, Van Bortel L, Boutouyrie P, Giannattasio C, Hayoz D, et al. Expert consensus document on arterial stiffness: methodological issues and clinical applications. Eur Heart J. 2006; 27:2588–2605. pmid:17000623
  3. 3. Westerhof N, Sipkema P, van den Bos GC, Elzinga G. Forward and backward waves in the arterial system. Cardiovasc Res. 1972; 6:648–656. pmid:4656472
  4. 4. Murgo JP, Westerhof N, Giolma JP, Altobelli SA. Aortic input impedance in normal man: relationship to pressure wave forms. Circulation. 1980; 62:105–116. pmid:7379273
  5. 5. Nichols WW, O'Rourke MF, Vlachopoulos C. McDonald's Blood Flow in Arteries. 6th ed. London, UK: Hodder Arnold; 2011.
  6. 6. Ben-Shlomo Y, Spears M, Boustred C, May M, Anderson SG, Benjamin EJ, et al. Aortic Pulse Wave Velocity Improves Cardiovascular Event Prediction: An Individual Participant Meta-Analysis of Prospective Observational Data From 17,635 Subjects. J Am Coll Cardiol. 2014; 63:636–646. pmid:24239664
  7. 7. Van Bortel LM, Laurent S, Boutouyrie P, Chowienczyk P, Cruickshank J, De Backer T, et al. Expert consensus document on the measurement of aortic stiffness in daily practice using carotid-femoral pulse wave velocity. J Hypertens. 2012; 30:445–448. pmid:22278144
  8. 8. Westerhof BE, Guelen I, Westerhof N, Karemaker JM, Avolio A. Quantification of Wave Reflection in the Human Aorta From Pressure Alone: A Proof of Principle. Hypertension. 2006; 48:595–601. pmid:16940207
  9. 9. Hametner B, Wassertheurer S, Kropf J, Mayer C, Holzinger A, Eber B, Weber T. Wave reflection quantification based on pressure waveforms alone—Methods, comparison, and clinical covariates. Comput Meth Prog Bio. 2013; 109:250–259.
  10. 10. Wang K-L, Cheng H-M, Sung S-H, Chuang S-Y, Li C-H, Spurgeon HA, et al. Wave Reflection and Arterial Stiffness in the Prediction of 15-Year All-Cause and Cardiovascular Mortalities: A Community-Based Study. Hypertension. 2010; 55:799–805. pmid:20065155
  11. 11. Weber T, Wassertheurer S, Rammer M, Haiden A, Hametner B, Eber B. Wave Reflections, Assessed With a Novel Method for Pulse Wave Separation, Are Associated With End-Organ Damage and Clinical Outcomes. Hypertension. 2012; 60:534–541. pmid:22585948
  12. 12. Chirinos JA, Kips JG, Jacobs DR, Brumback L, Duprez DA, Kronmal R, et al. Arterial Wave Reflections and Incident Cardiovascular Events and Heart Failure: The Multiethnic Study of Atherosclerosis. J Am Coll Cardiol. 2012; 60:2170–2177. pmid:23103044
  13. 13. Hametner B, Weber T, Mayer C, Kropf J, Wassertheurer S. Calculation of arterial characteristic impedance: a comparison using different blood flow models. Math Comput Model Dyn Syst. 2013; 19:319–330.
  14. 14. Butlin M, Avolio A. Noninvasive characterization of the effect of aortic impedance on left ventricular structure: a question of utility. J Hypertens. 2015; 33:63–65. pmid:25470120
  15. 15. Chirinos J, Segers P. Noninvasive Evaluation of Left Ventricular Afterload Part 1: Pressure and Flow Measurements and Basic Principles of Wave Conduction and Reflection. Hypertension. 2010; 56:555–562. pmid:20733089
  16. 16. Chirinos J, Segers P. Noninvasive Evaluation of Left Ventricular Afterload Part 2: Arterial Pressure-Flow and Pressure-Volume Relations in Humans. Hypertension. 2010; 56:563–570. pmid:20733088
  17. 17. Pucci G, Hametner B, Battista F, Wassertheurer S, Schillaci G. Pressure-independent relationship of aortic characteristic impedance with left ventricular mass and geometry in untreated hypertension. J Hypertens. 2015; 33:153–160. pmid:25269017
  18. 18. Wassertheurer S, Mayer C, Breitenecker F. Modeling arterial and left ventricular coupling for non-invasive measurements. Simul Model Pract Th. 2008; 16:988–997.
  19. 19. Rietzschel ER, De Buyzere ML, Bekaert S, Segers P, De Bacquer D, Cooman L, et al. Rationale, design, methods and baseline characteristics of the Asklepios Study. Eur J Cardiovasc Prev Rehabil. 2007; 14:179–191. pmid:17446795
  20. 20. Segers P, Rietzschel ER, De Buyzere ML, Vermeersch SJ, De Bacquer D, Van Bortel LM, et al., on behalf of the Asklepios investigators. Noninvasive (Input) Impedance, Pulse Wave Velocity, and Wave Reflection in Healthy Middle-Aged Men and Women. Hypertension. 2007; 49:1248–1255. pmid:17404183
  21. 21. Verbeke F, Segers P, Heireman S, Vanholder R, Verdonck P, Van Bortel LM. Noninvasive Assessment of Local Pulse Pressure: Importance of Brachial-to-Radial Pressure Amplification. Hypertension. 2005; 46:244–248. pmid:15911747
  22. 22. Hametner B, Wassertheurer S, Kropf J, Mayer C, Eber B, Weber T. Oscillometric estimation of aortic pulse wave velocity: comparison with intra-aortic catheter measurements. Blood Press Monit. 2013; 18:173–176. pmid:23571229
  23. 23. Luzardo L, Lujambio I, Sottolano M, da Rosa A, Thijs L, Noboa O, et al. 24-h ambulatory recording of aortic pulse wave velocity and central systolic augmentation: a feasibility study. Hypertens Res. 2012; 35:980–987. pmid:22622282
  24. 24. Parragh S, Hametner B, Bachler M, Weber T, Eber B, Wassertheurer S. Non-invasive wave reflection quantification in patients with reduced ejection fraction. Physiol. Meas. 2015; 36:179–190. pmid:25571922
  25. 25. Baumann M, Wassertheurer S, Suttmann Y, Burkhardt K, Heemann U. Aortic pulse wave velocity predicts mortality in chronic kidney disease stages 2–4. J Hypertens. 2014; 32:899–903. pmid:24609217
  26. 26. Kips JG, Rietzschel ER, De Buyzere ML, Westerhof BE, Gillebert TC, Van Bortel LM, Segers P. Evaluation of Noninvasive Methods to Assess Wave Reflection and Pulse Transit Time From the Pressure Waveform Alone. Hypertension. 2009; 53:142–149. pmid:19075098
  27. 27. Bland JM, Altman DG. Statistical methods for assessing agreement between two methods of clinical measurement. Lancet. 1986; 327:307–310.
  28. 28. Mitchell GF, Wang N, Palmisano JN, Larson MG, Hamburg NM, Vita JA, et al. Hemodynamic Correlates of Blood Pressure Across the Adult Age Spectrum: Noninvasive Evaluation in the Framingham Heart Study. Circulation. 2010; 122:1379–1386. pmid:20855656
  29. 29. Waddell TK, Dart AM, Gatzka CD, Cameron JD, Kingwell BA. Women exhibit a greater age-related increase in proximal aortic stiffness than men. J Hypertens. 2001; 19:2205–2212. pmid:11725165
  30. 30. Segers P, De Backer J, Devos D, Rabben SI, Gillebert TC, Van Bortel LM, et al. Aortic reflection coefficients and their association with global indexes of wave reflection in healthy controls and patients with Marfan's syndrome. Am J Physiol Heart Circ Physiol. 2006; 290:H2385–2392. pmid:16399860
  31. 31. Devos DG, Rietzschel E, Heyse C, Vandemaele P, Van Bortel L, Babin D, et al. MR pulse wave velocity increases with age faster in the thoracic aorta than in the abdominal aorta. J Magn Reson Imaging. 2015; 41(3):765–772. pmid:24615998
  32. 32. Vermeersch SJ, Rietzschel ER, De Buyzere ML, De Bacquer D, De Backer G, Van Bortel LM, et al., on behalf of the Asklepios investigators. Age and gender related patterns in carotid-femoral PWV and carotid and femoral stiffness in a large healthy, middle-aged population. J Hypertens. 2008; 26:1411–1419. pmid:18551018
  33. 33. Weber T, Wassertheurer S, Hametner B, Parragh S, Eber B. Non-invasive methods to assess pulse wave velocity: Comparison with the invasive gold-standard and relationship with organ damage. J Hypertens. 2015; 33(5):1023–31. pmid:25668350
  34. 34. Feistritzer H-J, Reinstadler SJ, Klug G, Kremser C, Seidner B, Esterhammer R, et al. Comparison of an Oscillometric Method with Cardiac Magnetic Resonance for the Analysis of Aortic Pulse Wave Velocity. PLoS ONE. 2015; 10(1):e0116862. pmid:25612307
  35. 35. Wilkinson IB, McEniery CM, Schillaci G, Boutouyrie P, Segers P, Donald A, Chowienczyk PJ. ARTERY Society guidelines for validation of non-invasive haemodynamic measurement devices: Part 1, arterial pulse wave velocity. Art Res. 2010; 4:34–40.
  36. 36. Mitchell GF, Parise H, Benjamin EJ, Larson MG, Keyes MJ, Vita JA, et al. Changes in Arterial Stiffness and Wave Reflection With Advancing Age in Healthy Men and Women: The Framingham Heart Study. Hypertension. 2004; 43:1239–1245. pmid:15123572
  37. 37. McEniery CM, Yasmin , Hall IR, Qasem A, Wilkinson IB, Cockroft JR. Normal Vascular Aging: Differential Effects on Wave Reflection and Aortic Pulse Wave Velocity. J Am Coll Cardiol. 2005; 46:1753–1760. pmid:16256881
  38. 38. Smulyan H, Asmar RG, Rudnicki A, London GM, Safar ME. Comparative Effects of Aging in Men and Women on the Properties of the Arterial Tree. J Am Coll Cardiol. 2001; 37:1374–1380. pmid:11300449
  39. 39. Tomiyama H, Yamashina A, Arai T, Hirose K, Koji Y, Chikamori T, et al. Influences of age and gender on results of noninvasive brachial-ankle pulse wave velocity measurement—a survey of 12517 subjects. Atherosclerosis. 2003; 166:303–309. pmid:12535743